paper_id
stringlengths 9
16
| version
stringclasses 26
values | yymm
stringclasses 311
values | created
timestamp[s] | title
stringlengths 6
335
| secondary_subfield
sequencelengths 1
8
| abstract
stringlengths 25
3.93k
| primary_subfield
stringclasses 124
values | field
stringclasses 20
values | fulltext
stringlengths 0
2.84M
|
---|---|---|---|---|---|---|---|---|---|
1008.3478 | 3 | 1008 | 2011-08-30T13:31:10 | Spatial clustering of interacting bugs: Levy flights versus Gaussian jumps | [
"physics.bio-ph",
"cond-mat.stat-mech"
] | A biological competition model where the individuals of the same species perform a two-dimensional Markovian continuous-time random walk and undergo reproduction and death is studied. The competition is introduced through the assumption that the reproduction rate depends on the crowding in the neighborhood. The spatial dynamics corresponds either to normal diffusion characterized by Gaussian jumps or to superdiffusion characterized by L\'evy flights. It is observed that in both cases periodic patterns occur for appropriate parameters of the model, indicating that the general macroscopic collective behavior of the system is more strongly influenced by the competition for the resources than by the type of spatial dynamics. However, some differences arise that are discussed. | physics.bio-ph | physics | epl draft
1
1
0
2
g
u
A
0
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
3
v
8
7
4
3
.
8
0
0
1
:
v
i
X
r
a
Spatial clustering of interacting bugs:
L´evy flights versus Gaussian jumps
E. Heinsalu1,2, E. Hern´andez-Garc´ıa1 and C. L´opez1
1 IFISC, Instituto de F´ısica Interdisciplinar y Sistemas Complejos (CSIC-UIB), E-07122 Palma de Mallorca, Spain
2 National Institute of Chemical Physics and Biophysics, Ravala 10, Tallinn 15042, Estonia
PACS 02.50.Ey -- Stochastic processes
PACS 05.40.-a -- Fluctuation phenomena, random processes, noise, and Brownian motion
PACS 05.40.Fb -- Random walks and Levy flights
Abstract. - A biological competition model where the individuals of the same species perform
a two-dimensional Markovian continuous-time random walk and undergo reproduction and death
is studied. The competition is introduced through the assumption that the reproduction rate de-
pends on the crowding in the neighborhood. The spatial dynamics corresponds either to normal
diffusion characterized by Gaussian jumps or to superdiffusion characterized by L´evy flights. It
is observed that in both cases periodic patterns occur for appropriate parameters of the model,
indicating that the general macroscopic collective behavior of the system is more strongly influ-
enced by the competition for the resources than by the type of spatial dynamics. However, some
differences arise that are discussed.
Introduction. -- Interacting particle systems help to
model and understand various problems in many diverse
fields. In biological contexts they are particularly impor-
tant to study aggregation phenomena of individuals. Fish
schools, insect swarms, bacterial patterns, bird flocks and
patchy plankton structures are just a few examples reveal-
ing the ubiquity and fundamental importance of organism
aggregates.
An attempt to address a simple mechanism giving rise
to the clustering of particles (with emphasis on plankton
patchiness) was made within a Brownian bug model [1]
(see also refs. [2, 3]). The model consists of an ensemble
of particles (bugs), each one dying or reproducing with
a given probability and undergoing Brownian motion. If
the diffusivity of the particles is low enough, macroscopic
spatial clustering occurs, since a newborn is located close
to the parent but particles can die anywhere. If diffusivity
is large, particles perform more extended walks and the
region that was left empty due to the death of a particle, is
occupied fast by some other particle. Similar results were
obtained in refs. [4 -- 7] where lattice-models were studied.
The basic Brownian bug model lacks any interaction
between the particles.
In a more realistic model (inter-
acting Brownian bug model) the inter-particle interaction
was taken into account assuming that the birth and death
of individuals depend on the number of other bugs in
the neighborhood [8 -- 13]. For appropriate parameters, a
salient property of that model is the formation of spa-
tially periodic clustering of bugs [8]. For large diffusion
the clusters become blurred and the periodic pattern is
replaced by a more uniform distribution of bugs. Impor-
tantly, whereas the positive correlations leading to clus-
tering in the non-interacting bug case arise from the re-
productive correlations, i.e., from the fact that offspring
is born at the same location of parent [5, 6], the periodic
arrangement of clusters in the system of interacting bugs
is a consequence of the competitive interaction and has a
spatial scale determined by the interaction range R [8].
At the same time it has been observed that many liv-
ing organisms move consistently with L´evy flight behav-
ior [14 -- 18]. In particular, the motion of some bacteria is
found to be described by L´evy statistics [19, 20], as well
as the movement of spider monkeys in search of food [21].
The L´evy type of motion has been shown to be advanta-
geous with respect to standard Brownian motion in some
searching strategies involving foraging [18], or in order to
enhance encounter rates at low densities [22]. The main
reason for this resides in the occurrence of occasional long
jumps.
However, the impact of L´evy-type diffusion on the prop-
erties of organism aggregates has not received much atten-
tion thus far. In the present paper we investigate in the
p-1
E. Heinsalu et al.
context of a simple interacting bug model, continuous in
space, how the occurrence of long jumps in the motion
of the individuals influences the collective behavior (c.f.
ref. [23]). Similarities and differences between the inter-
acting Brownian and L´evy bug systems will be highlighted.
Model and numerical algorithm. -- We consider a
system consisting initially of N0 pointlike particles, placed
randomly in a two-dimensional L× L square domain with
periodic boundary conditions. After the random time τ ,
a particle i, chosen randomly among all the N (t) bugs in
the system at the present time t, undergoes one of the two
events: it either reproduces or disappears. For the birth
and death rates of the i-th particle we assume [8],
ri
ri
b = max(cid:0)0, rb0 − αN i
d = max(cid:0)0, rd0 + βN i
R(cid:1) ,
R(cid:1) .
(1)
(2)
Here N i
R is the number of particles which are at a distance
smaller than R from particle i, the parameters rb0 and rd0
are the zero-density birth and death rates, and the pa-
rameters α and β determine how ri
d depend on the
neighborhood; the function max() enforces the positivity
of the rates. Such a choice for the reproduction and death
rates introduces an interaction between the bugs. For the
random times τ an exponential probability density with
the complementary cumulative distribution
b and ri
p(τ ) = exp(−τ /τ )
(3)
and a characteristic time hτi = τ = R−1
the time-scale parameter and
tot is chosen; τ is
N
Rtot =
(ri
b + ri
d)
Xi=1
(4)
is the total rate of all the N ≡ N (t) particles.
In the
case of reproduction, the new bug is located at the same
position (xi, yi) as the parent particle i. After the de-
mographic event, i.e., after each random time τ , all the
particles perform a jump of random length ℓ in a random
direction characterized by an angle uniformly distributed
on (0, 2π) (ℓ and the direction of the jump are different
for each particle). The new present time is t + τ and the
process is repeated again, until the final simulation time
is reached [13].
In order to simulate the system where the particles un-
dergo normal diffusion, a Gaussian jump-length probabil-
ity density function is used,
ϕ(ℓ) = (cid:16)ℓ√2π(cid:17)−1
exph−ℓ2/(2ℓ2)i ,
(5)
with variance hℓ2i = ℓ2; ℓ is the space-scale parameter.
Since we draw the angle specifying the direction of the
jump from the interval (0, 2π), we can neglect the sign
of ℓ. Note that the random walk defined in this way is
not exactly the same as the one in which the walker per-
forms jumps extracted from a two-dimensional Gaussian
distribution, but it also leads to normal diffusion and al-
lows a more direct comparison with the L´evy case. The
corresponding diffusion coefficient is
κ = hℓ2i/(2hτi) ,
(6)
which expressed through the space- and time-scale pa-
rameters reads, κ = ℓ2/(2τ ). As we choose to fix the
value of κ, then the space-scale parameter is determined
by ℓ = √2κτ = p2κ/Rtot. Note that because the number
of particles is changing in time also the quantity Rtot is
changing in time. However, as in this paper we investigate
asymptotic statistically steady states, the number of par-
ticles weakly fluctuates around a well-defined mean value.
Thus, the space-scale parameter ℓ is fluctuating in time,
but this time dependence is not affecting qualitatively the
dynamics of the system.
In order to simulate the system where the particles un-
dergo superdiffusion one can use a symmetric L´evy stable
probability density function for the jump size, which be-
haves asymptotically as [14, 17]
ϕµ(ℓ) ≈ ℓµℓ−µ−1 ,
ℓ → ±∞ (ℓ ≫ ℓ) ,
(7)
with the L´evy index 0 < µ < 2. For all L´evy stable
probability density functions with µ < 2 the variance di-
verges, hℓ2i = ∞, leading to the occurrence of extremely
long jumps, and typical trajectories are self-similar, on all
scales showing clusters of shorter jumps intersparsed by
long excursions. For 0 < a < µ < 2 fractional moments
converge, hℓai < ∞. Therefore, for the L´evy index in
the range 1 < µ < 2 the value of hℓi is finite, whereas
for 0 < µ ≤ 1 it diverges. The complementary cumulative
distribution corresponding to (7) is
Pµ(ℓ) ≈ µ−1(ℓ/ℓ)−µ ,
ℓ → ±∞ .
(8)
As a simple form of complementary cumulative distribu-
tion function, which is normalizable and behaves asymp-
totically as (8), we use
Pµ(ℓ) = (1 + b1/µℓ/ℓ)−µ ,
(9)
where b = [Γ(1 − µ/2)Γ(µ/2)]/Γ(µ). As before, we can
neglect the sign of ℓ since the direction of the jump is
assigned by drawing an angle on (0, 2π). Now the diffusion
coefficient (6) is infinite, but one can define a generalized
diffusion coefficient in terms of the scales ℓ and τ as [14,17]
κµ = ℓµ/(2τ ) .
(10)
Therefore, in the case of the L´evy flights, when fixing the
value of κµ, the space-scale parameter is, ℓ = (2κµ τ )1/µ =
(2κµ/Rtot)1/µ.
The model described can be interpreted in the follow-
ing way: during a time interval τ each individual moves
on average a distance ℓ in some direction but only one
of them reproduces or dies. For positive values of α and
β, the more neighbors a particle has within the radius
p-2
Spatial clustering of interacting bugs: L´evy flights versus Gaussian jumps
(a)
y
(b)
y
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1
x
0
0
0.2
0.4
0.6
0.8
1
x
Interacting Brownian bugs (a) versus interacting L´evy
Fig. 1:
bugs (b): spatial configuration of particles in the statistically
steady state. The parameters are: rb0 = 1, rd0 = 0.1, α = 0.02,
β = 0, and R = 0.1 (see also text). In the case of Brownian
bugs the diffusion coefficient is κ = 10−5. In the case of L´evy
bugs the value of L´evy index is µ = 1 and the anomalous
diffusion coefficient is κµ = 56 × 10−5. The average number of
particles in the two systems is approximately the same, hN i =
2555 and hN i = 2565, respectively.
R, the smaller is the probability of reproduction and the
larger is the probability that the bug does not survive,
e.g., due to competition for resources. The periodicity of
the simulation domain represents the fact that particles
can leave and enter the observation area. In the case of
long jumps it represents the situation where there is one
particle that leaves the domain and another one that ar-
rives from somewhere, perhaps from far away, and its po-
sition is basically random in the domain. In principle, one
could also simulate the system so that after the inter-event
time τ a randomly chosen particle undergoes one of the
three events: reproduction, death, or jump. However, the
results obtained are the same as following the procedure
described above; one would just have one more parameter,
the jump rate, and different numerical values for the other
parameters.
Results. -- In the following we set β = 0 in eq. (2)
and α = 0.02 in eq. (1), i.e., the probability of death is
constant and the same for all the particles while the prob-
ability for the reproduction depends on how crowded the
environment is. Also, we use the values rb0 = 1, rd0 = 0.1,
L = 1, and R = 0.1 in all the figures presented in the
current paper. For β = 0 the critical number of neigh-
bors, N ∗
R, for which death and reproduction are equally
probable for particle i, is determined by
N ∗
R = ∆bd/α ,
(11)
where ∆bd = rb0 − rd0. If N i
R < N ∗
R it is more probable
that particle i reproduces and if N i
R > N ∗
R death is more
likely. In the following we fix the value of the L´evy index
to µ = 1. The detailed discussion of the influence of the
birth and death rates as well as the L´evy index on the
results is outside the scope of the current paper and will
be presented elsewhere.
In the case of interacting Brownian bugs,
for small
enough κ and large enough ∆bd, the occurrence of pe-
riodic patterns has been observed previously [8, 10].
In
the statistically steady state clusters form and arrange
in a hexagonal lattice (see fig. 1-a). For large values of
the diffusion coefficient the periodic pattern is replaced
by an almost homogeneous distribution of particles (see
also fig. 3-a).
In the case of L´evy flights, since the dif-
fusion coefficient (6) is infinite, one could expect that for
the interacting L´evy bugs the spatial distribution will not
reveal a periodic pattern. However, as can be seen from
fig. 1-b, this is not the case. The reason for the diver-
gence of the diffusion coefficient in the L´evy case is in the
statistical weight of the large jumps. These large jumps
have some influence in the characteristics of the patterns
formed, but the large-scale structure is ruled mainly by
the interactions between particles.
For the interacting bug system with Gaussian jumps and
moderate diffusion the appearance of periodic clustering
is well captured by combining the effects of diffusion and
interaction in a mean-field approach [8]. Following the
steps in ref. [8] one obtains the following equation as a
mean-field approximation to the dynamics of the density of
particles ρ(x, t) in the interacting L´evy bug model (β = 0):
∂ρ(x, t)
∂t
= ρ(x, t)(cid:18)∆bd − αZD
dyρ(y, t)(cid:19) + κµ∇γρ(x, t) .
(12)
The integration domain D is the set of points within a dis-
tance smaller than R from x: x − y < R. The operator
∇γ, with γ = µ1, is the fractional diffusion operator asso-
ciated to the L´evy flights of exponent µ ∈ (0, 2) [14, 17].
It reduces to the standard diffusion operator for µ > 2. In
this mean-field description the first term accounts for the
net growth of the population, the second one takes into
1This expression was incorrect in the paper originally published
in EPL 92, 40011 (2010); it was corrected to the present form in the
Erratum in EPL 95, 69902 (2011).
p-3
E. Heinsalu et al.
100
10
1
0.1
)
r
(
g
Gaussian
Levy
0.01
0.1
r
0.5
Fig. 2: Comparison of the radial distribution functions for
the systems with Gaussian jumps and with L´evy flights. The
parameters are as in fig. 1.
account the non-local contribution associated to the sat-
uration due to the interactions within a distance R, and
the third term describes the spatial diffusion of particles.
Equation (12) differs from the mean-field approximation
derived in ref. [8] only in the third term where the stan-
dard diffusion operator is replaced by the fractional diffu-
sion operator and the diffusion coefficient by the general-
ized diffusion coefficient. Similar descriptions of reaction-
diffusion systems with a fractional diffusion term can be
consulted, e.g., in refs. [14, 15, 24, 25].
Equation (12) has always the uniform solution
ρ(x, t) = ρ0 = ∆bd/(απR2) ,
(13)
which turns out to become unstable for small κµ and/or
large ∆bd (what is small or large depends also on the value
of µ). This can be seen by introducing the ansatz ρ(x, t) =
ρ0 + δρ(x, t) in eq. (12) and linearizing in δρ(x, t). The
result is
δρ(x, t) ∼ exp(ik · x + λ(k)t) ,
with
λ(k) = −κµkγ − 2∆bdJ1(kR)/(kR) ;
(14)
(15)
J1 is a Bessel function. An instability of the uniform solu-
tion will occur if the sign of λ(k) changes from negative
to positive at some value of k. This will occur first for
the critical wavenumber kc for which λ(k) is maximum.
The instability will develop in a periodic pattern which, at
least close enough to the instability, will have a periodic-
ity δ = 2π/kc. Introducing a dimensionless wavenumber
q ≡ kR and growth rate Λ = Rγλ/κµ, eq. (15) reads,
Λ(q) = −qγ − νJ1(q)/q .
(16)
The latter equation shows that, within this mean-field de-
scription, the relevant parameters are γ (or µ) and the
dimensionless quantity ν ≡ 2Rγ∆ab/κµ.
Imposing the
condition Λ(q) = 0, corresponding to the change of sign in
(a) κ = 6⋅10-6
10-5
3⋅10-5
10-4
(b)
κµ = 3⋅10-4
6⋅10-4
10-3
5⋅10-3
100
10
1
0.1
10
1
)
r
(
g
)
r
(
g
0.01
0.1
r
0.5
Fig. 3: Radial distribution function for the system with (a)
Gaussian jumps and (b) L´evy jumps for several values of the
diffusion and generalized diffusion coefficient, respectively. The
other parameters are as in fig. 1 (see also text).
the growth rate, and Λ′(q) = 0, corresponding to the max-
imum of the growth rate, one can find the critical value
of ν, νc = −qγ+1
/J1(qc), and the equation for the critical
wavenumber:
c
qcJ2(qc)/J1(qc) = −γ .
(17)
Numerical solution of the latter equation provides qc, and
therefore δ, as a function of γ (or µ), and that δ is propor-
tional to R. In order to make a comparison with the results
from numerical simulations of the system, one should keep
in mind that the periodic boundary conditions will change
qc to the closest number of the form (2πR/L)√n2 + m2
with n and m integers. For µ = 1 the value of δ ob-
tained from the mean-field approximation is 0.128036 and
for µ = 2 (Gaussian case) it is 0.131306, i.e., the periodic-
ity of the pattern is in both cases of the order of R = 0.1.
To extract the periodicity of the pattern from the nu-
merical simulations of the system, it is useful to study
the radial distribution function g(r).
It describes how
the density varies with the distance from a given parti-
cle respect to the one expected from a uniform distribu-
tion, giving thus additional information about the distri-
bution of bugs [26]. It is computed in the standard way,
i.e., by counting all particles, dn, at a distance between
p-4
Spatial clustering of interacting bugs: L´evy flights versus Gaussian jumps
Gaussian
Levy
ρ
100
80
60
40
20
0
0
10
20
30
40
50
Nc
Fig. 4: The distribution of cluster sizes for the interacting
Brownian bug and L´evy bug models. All parameters have the
same values as in figs. 1 and 2.
r and r + dr from the target particle, using the formula
dn = (N/L2)g(r)2πrdr, and averaging over all particles
and different long times. In fig. 2 the comparison of the ra-
dial distribution functions for the Brownian and L´evy bugs
(for the same systems as in fig. 1) is depicted. Figure 3-a
shows the behavior of g(r) in the case of Brownian bugs
and in fig. 3-b in the case of L´evy bugs for various values
of diffusion coefficient and generalized diffusion coefficient,
respectively. The second maximum of g(r), indicating the
periodicity of the pattern, appears in fig. 3-a (in the sys-
tems of Brownian bugs), up to κ = 10−5, at 0.13125 and in
fig. 3-b (in the L´evy case), up to κµ = 10−3, at 0.12875, be-
ing in good agreement with the results obtained from the
mean-fieled approximation, and becoming slightly larger
only at larger values of κ and κµ. The anomalous expo-
nent µ has only a very light influence on the periodicity
of the pattern, which is of the order of R (see also fig. 2
where the slight shift can be can be noticed).
Figure 3 shows also that as κ or κµ increases, the first
peak of g(r) gets lower and wider: the total number of
particles in the system decreases and the clusters become
more spread. For large values of diffusivity (κ or κµ),
the oscillations of g(r) smooth out, indicating that the
periodic pattern becomes replaced by a more homogeneous
distribution of bugs.
As a difference compared to the case of interacting
Brownian bugs, we observe that now, even at small val-
ues of κµ, there are many solitary particles appearing for
short time periods in the space between the periodically
arranged clusters, due to the large jumps, c.f. figs. 1-a
and 1-b. This is even better illustrated by fig. 4 where
the probability distributions of the cluster sizes for the
systems with Gaussian and L´evy jumps are depicted (for
the same systems as in fig. 1, i.e., for the given param-
eters the values of κ and κµ have been chosen such that
the average number of the particles in the two systems is
approximately equal). The clusters are defined using the
nearest neighbor clustering method [27], with a threshold
(a)
Gaussian
Levy
8⋅104
6⋅104
4⋅104
ρ
2⋅104
(b)
0
104
103
102
ρ
101
-0.06
-0.03
0.03
0.06
0
x
Fig. 5: Cross-section of the two-dimensional particle density
of the average cluster in (a) linear and (b) semi-log scale. Com-
parison between the interacting Brownian and L´evy bug mod-
els. The values of the parameters are the same as in figs. 1,
2, and 4. The continuous lines correspond to the fitting with
Gaussian functions.
value 0.025 to define neighbors. As one can see, for the
system of interacting Brownian bugs the most probable
size of large clusters forming the periodic pattern is 41
particles whereas for interacting L´evy bugs it is 36 (the
critical number of neighbors is in both cases N ∗
R = 45).
A noticeable difference occurs in the proportion of single-
particle clusters: the peak of the distribution at the value
characterizing isolated particles (Nc = 1) is much higher
in the L´evy flight case. Returning to fig. 2, which refers
to the same systems as in figs. 1 and 4, it is to be noticed
that the minimum between the first and the second peak is
much lower for the system of Brownian bugs. This result
is consistent with fig. 4: in the system with L´evy flights,
the occurrence of many single particles between the clus-
ters forming the periodic pattern produces a higher inter-
cluster density is observed.
Finally, in fig. 5 the cross-section of the two-dimensional
particle density of the average cluster is depicted for the in-
teracting Brownian and L´evy bug systems (same systems
as in fig. 1). The average cluster is obtained by setting
the origin at the center of mass of each cluster forming the
periodic pattern and averaging over all the clusters in the
p-5
E. Heinsalu et al.
simulation area and over time. In the case of Brownian as
well as of L´evy bugs, the central part of the average clus-
ter, where most of the particles are concentrated, is well
fitted by a Gaussian function (continuous curve). How-
ever, fig. 5-b reveals the difference in the way the particle
density decreases in the two systems when moving away
from the center of mass of the cluster. Though even in
the case of the Brownian bugs the tail of the average clus-
ter is not really well described by a Gaussian function, a
Gaussian decay provides a good approximation. Instead,
in the case of the L´evy bugs the tail of the average clus-
ter decays slower. Since the average cluster is calculated
from individual clusters which are separated by a distance
of order R, one cannot properly estimate an asymptotic
decay, but the decay in the L´evy case seems close to ex-
ponential. In any case, it is distinctively faster than the
power-law behavior which would be exhibited by clusters
of non-interacting particles moving purely by L´evy flights.
The existence of the interaction range R introduces a cut-
off distance which makes the long-jumps characteristic of
L´evy flights not so relevant to determine the large-scale
properties of the spatial particle configurations.
Conclusion. -- In the present paper we have studied
a simple birth-death model with competition among the
individuals of the same species.
In particular, we have
investigated how the superdiffusive motion of the individ-
uals characterized by L´evy flights influences the collective
behavior. We have observed that the appearance of peri-
odic clustering known from previous studies of the inter-
acting Brownian bug model takes place also in the inter-
acting L´evy bug model for appropriate parameters. This
is against the expectation that particles performing L´evy
flights cannot give rise to space-periodic clustering since
their diffusion coefficient is infinite, for which the interact-
ing Brownian bug model is known to reveal no periodic
pattern. However, as a difference we have observed that,
in the interacting L´evy bug model, due to the long jumps,
there are many single particles between the clusters, lead-
ing also to the differences in the particle density profiles
of the average cluster or in the short-distance behavior of
the radial distribution function. From this one can con-
clude that the large-scale collective behavior of the system
is much more strongly influenced by the competitive in-
teraction than by the type of spatial motion performed by
the bugs. We have also verified that the mean-field ap-
proximation (12) is proper to describe the periodicity in
the interacting L´evy bug model.
As a final remark, let us mention that though the model
studied in the present article describes rather living organ-
isms, such as animals or bacteria, non-local interactions
and L´evy flights are important also in plant ecology for
the development of vegetation patterns [28], due to the
competition for the resources and because the seed dis-
persal is often described better by L´evy flights than by
Gaussian jumps [29].
∗ ∗ ∗
This work has been supported by the targeted financ-
ing project SF0690030s09, Estonian Science Foundation
through grant no.
7466, by the Balearic Government
(E.H.), and by Spanish MICINN and FEDER through
project FISICOS (FIS2007-60327).
REFERENCES
[1] Young W. R., Roberts A. J. and Stuhne G., Nature
, 412 (2001) 328.
[2] Zhang Y.-C., Serva M. and Polikarpov M., J. Stat.
Phys. , 58 (1990) 849.
[3] Felsenstein J., Am. Nat. , 109 (1975) 359.
[4] Houchmandzadeh B., Phys. Rev. E , 66 (2002) 052901.
[5] Houchmandzadeh B., Phys. Rev. Lett. , 101 (2008)
078103.
[6] Houchmandzadeh B., Phys. Rev. E , 80 (2009) 051920.
[7] Paessens M. and Schutz G., J. Phys. A , 37 (2004)
4709.
[8] Hern´andez-Garc´ıa E. and L´opez C., Phys. Rev. E ,
70 (2004) 016216.
[9] Hern´andez-Garcia E. and L´opez C., J. Phys.: Con-
dens. Matter , 17 (2005) S4263.
[10] L´opez C. and Hern´andez-Garcia E., Physica D , 199
(2004) 223.
[11] Bolkel B. M. and Pacala S. W., Am. Nat. , 153 (1999)
575.
[12] Martin A. P., J. Theor. Biol. , 230 (2004) 343.
[13] Birch D. A. and Young W. R., Theor. Pop. Biol. , 70
(2006) 26.
[14] Metzler R. and Klafter J., Phys. Rep. , 339 (2000)
1.
[15] Metzler R. and Klafter J., J. Phys. A: Math. Gen. ,
37 (2004) R161.
[16] Dieterich P., Klages R., Preuss R. and Schwab A.,
Proc. Natl. Acad. Sci. USA , 105 (2008) 459.
[17] Klages R., Radons G. and Sokolov I. M., Anomalous
Transport: Foundations and Applications (Wiley-VCH)
2008.
[18] Viswanathan G. M., Raposo E. P. and da Luz
M. G. E., Phys. Life Rev. , 5 (2008) 133.
[19] Levandowsky M., White B. S. and Schuster F. L.,
Acta Protozool. , 36 (1997) 237.
[20] Nossal R., J. Stat. Phys. , 30 (1983) 391.
[21] Ramos-Fern´andez G., Mateos J. L., Miramontes
O., Cocho G., Larralde H. and Ayala-Orozco B.,
Behav. Ecol. Sociobiol. , 55 (2004) 223.
[22] Bartumeus F., Peters F., Pueyo S., Marras´e C. and
Catal´an J., Proc. Natl. Acad. Sci. USA , 100 (2003)
12771.
[23] Durang X. and Henkel M., J. Phys. A , 42 (2009)
395004.
[24] Baeumer B., Kov´acs M. and Meerschaert M., Bull.
Math. Biol. , 69 (2007) 2281.
[25] del Castillo-Negrete D., Carreras B. A. and
Lynch V. E., Phys. Rev. Lett. , 91 (2003) 018302.
[26] Ramos F., L´opez C., Hern´andez-Garc´ıa E. and
Munoz M. A., Phys. Rev. E , 77 (2008) 021102.
p-6
Spatial clustering of interacting bugs: L´evy flights versus Gaussian jumps
[27] Florek K., Lukaszewicz J., Steinhaus H. and
Zubrzycki S., Colloquium Mathematicum , 2 (1951) 282.
[28] Borgogno F., D'Odorico P., Laio F. and Ridolfi L.,
Rev. Geophys. , 47 (2009) RG1005.
[29] Clark J. S., Silman M., Kern R., Macklin E. and
HilleRisLambers J., Ecology , 80 (1999) 1475.
p-7
|
1604.06366 | 1 | 1604 | 2016-04-21T16:02:03 | Molecular dynamics study of accelerated ion-induced shock waves in biological media | [
"physics.bio-ph",
"cond-mat.soft"
] | We present a molecular dynamics study of the effects of carbon- and iron-ion induced shock waves in DNA duplexes in liquid water. We use the CHARMM force field implemented within the MBN Explorer simulation package to optimize and equilibrate DNA duplexes in liquid water boxes of different sizes and shapes. The translational and vibrational degrees of freedom of water molecules are excited according to the energy deposited by the ions and the subsequent shock waves in liquid water are simulated. The pressure waves generated are studied and compared with an analytical hydrodynamics model which serves as a benchmark for evaluating the suitability of the simulation boxes. The energy deposition in the DNA backbone bonds is also monitored as an estimation of biological damage, something which lies beyond the possibilities of the analytical model. | physics.bio-ph | physics | EPJ manuscript No.
(will be inserted by the editor)
6
1
0
2
r
p
A
1
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
6
6
3
6
0
.
4
0
6
1
:
v
i
X
r
a
Molecular dynamics study of accelerated ion-induced
shock waves in biological media
Pablo de Vera1,2,3,a, Nigel J. Mason2, Fred J. Currell1, and Andrey V. Solov'yov3,b
1 School of Mathematics and Physics, Queen's University Belfast, BT7 1NN Belfast, Northern Ireland, United Kingdom
2 Department of Physical Sciences, The Open University, Walton Hall, MK7 6AA Milton Keynes, England, United Kingdom
3 MBN Research Center, Altenhoferallee 3, 60438 Frankfurt am Main, Germany
Received: date / Revised version: date
Abstract. We present a molecular dynamics study of the effects of carbon- and iron-ion induced shock waves
in DNA duplexes in liquid water. We use the CHARMM force field implemented within the MBN Explorer
simulation package to optimize and equilibrate DNA duplexes in liquid water boxes of different sizes and
shapes. The translational and vibrational degrees of freedom of water molecules are excited according
to the energy deposited by the ions and the subsequent shock waves in liquid water are simulated. The
pressure waves generated are studied and compared with an analytical hydrodynamics model which serves
as a benchmark for evaluating the suitability of the simulation boxes. The energy deposition in the DNA
backbone bonds is also monitored as an estimation of biological damage, something which lies beyond the
possibilities of the analytical model.
1 Introduction
Collisional phenomena between fast ions and biomolecules
is a topic of major interest since it is necessary to un-
derstand the mechanisms of such processes if we are to
develop ion beam cancer therapy (IBCT). In this ther-
apy technique, energetic protons or heavier ions are used
clinically to treat deeply seated tumors [1]. From a macro-
scopic point of view one of the working principles of IBCT
is the Bragg peak, a sharp maximum in the depth-dose
curve at the end of the energetic ion trajectories (con-
trary to photon or electron beams which have a quite
broad energy deposition profile) that maximizes energy
deposition in the tumor while sparing surrounding healthy
tissues. However, it is well known that the effectiveness
of IBCT relies on nanoscopic phenomena rather than on
macroscopic characteristics [2,3], the former being directly
related to atomic collisions with biomolecules. Indeed, a
given dose deposited by ions presents a much larger cell
killing probability than the same dose deposited by pho-
tons. This increased relative biological effectiveness is due
to the large energy deposited around ion tracks on the
nanoscale, giving place to an increase in the clustering
of damaging events in biomolecules, especially in nuclear
DNA, which makes the repair processes less effective [4].
The fundamental aspects of the problem are also of
great interest. The irradiation with ions involves new phys-
ical phenomena which are not always considered properly
a Corresponding author: [email protected]
b On leave from A. F. Ioffe Physical Technical Institute,
194021 St. Petersburg, Russian Federation
(or considered at all) in biophysical models. In fact, IBCT
is a complex problem, involving many different space, en-
ergy, and time scales, ranging from the transport of en-
ergetic ions in macroscopic tissues, the production of sec-
ondary electrons and radicals that can propagate on the
nano- and microscale (molecular and cellular levels, re-
spectively) and their interaction with biomolecules on the
nanometer scale leading to the final biological outcomes,
noticeable in larger space and time scales [3]. More sig-
nificantly, all the physico-chemical processes occurring at
the molecular level make it necessary to deviate from a
simple energy deposition scheme. On these space scales
not only energy deposition events are of importance, since
the way in which this energy promotes different kind of
processes can significantly affect the final effects. While
many processes are fairly well known, such as the electron
production and propagation or the generation of free rad-
icals, new interactions are being discovered, such as the
dissociative electron attachment, a mechanism by which
very low energy electrons (with energies even below the
ionization threshold) can fragment biomolecules [5].
In this context another damage mechanism has been
theoretically predicted: ion-induced shock waves on the
nanometer scale [6,7,8]. Ion beams can deposit large
amounts of energy per unit path length (a carbon ion
in the Bragg peak region deposits 900 eV/nm), and the
major part of this energy is used to eject secondary elec-
trons of very low energies, below 50 eV [3,9,10]. Most of
such electrons transfer their energy to electronic excita-
tions of the medium in less than a nanometer and the
time scale in which this energy loss occurs is very short,
2
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
of a few femtoseconds [11]. These times are very short
in comparison with the mechanism capable of dissipating
this energy, the electron-phonon coupling, which occurs in
the sub-picosecond scale [11]. This situation results in a
large heating of the medium in nanocylinders around the
ion tracks, providing the conditions for a violent explo-
sion of these "hot cylinders", a mechanism we refer to as
ion-induced shock waves.
This shock wave effect was first predicted in terms of a
hydrodynamics model where it was shown that pressures
up to tens of GPa can be produced around ion tracks
[7]. However, this is not enough for predicting biomolecu-
lar damage and subsequently molecular dynamics simula-
tions were used to show how these conditions are sufficient
to produce bond breaking in nucleosomes [6]. Moreover
since the shock waves travel at high velocity they can
propagate secondary species (i.e., free radicals, solvated
electrons) much faster than the diffusion mechanism. All
these dynamical and thermo-mechanical effects can dras-
tically change the physico-chemical environment to which
biomolecules are exposed under irradiation.
A proper understanding of the characteristics of shock
waves requires more systematic studies, especially by the
use of molecular dynamics, a technique that can assess
their properties and consequences in more detail. Such
studies will enhance the understanding of the biological
role of ion-induced shock waves, as well as help in the
design of possible experiments to verify their existence.
In the present paper we report molecular dynamics sim-
ulations to study the main features of ion-induced shock
waves and their effects on DNA by means of the MBN
Explorer software [13]. We focus the work on shock waves
produced around carbon (one of the most promising pro-
jectiles used in IBCT) and iron ion tracks in the Bragg
peak region. We study the dependence of several quanti-
ties, namely pressure waves and energy deposition in DNA
bonds, on the size and characteristics of the system, in
order to establish the proper features of the simulation
box for future more systematic studies. These results are
compared with the analytical hydrodynamics model [7] as
well as to previous simulation results [6] to benchmark
the simulations. As a probe for biodamage we will use
short DNA segments. This is because such short DNA
duplexes can be simulated more straightforwardly than
nucleosomes, so they are more convenient for systematic
studies. Also, since the effects of shock waves are more
noticeable in short distances, a short DNA strand should
be enough for evaluating biomolecular damage, in a sim-
ilar fashion to the effects of secondary electrons on DNA
studied previously [3].
The methodology of the work is explained in section 2,
where the hydrodynamics model (subsection 2.1) and the
molecular dynamics procedure (subsection 2.2) are dis-
cussed. The results of the simulations are presented in sec-
tion 3, where the pressure generated by the shock waves
and their effects in the DNA duplex are studied. The final
conclusions and remarks are given in section 4.
2 Methods
Ion-induced nanoscopic shock waves were first described
in terms of an analytical hydrodynamics model that was
used for the first evaluation of their characteristics [7].
Even though it allows calculation of the basic physical
features, such as the pressure generated during the shock
wave, it does not allow the evaluation of the characteristics
related to biological effects, such as the damage of DNA
molecules, for which an atomic level description is needed.
This kind of analysis can be performed through molecular
dynamics simulations [6]. However, the analytical hydro-
dynamics model serves as a good benchmark for the molec-
ular dynamics simulations. Some relevant features of the
analytical hydrodynamics model are reviewed in subsec-
tion 2.1. The molecular dynamics simulations technique is
reviewed in subsection 2.2. The choice of the biomolecu-
lar probe, as well as its setting up for the simulations, is
explained in subsection 2.2.1, and the setting up of the ini-
tial conditions for the simulation of the ion-induced shock
wave is described in subsection 2.2.2.
2.1 Hydrodynamics model
This model was adapted from a classical hydrodynam-
ics treatment of the self-similar flow of liquid water and
heat transfer [12] and applied to the specific situation of
the energy delivered around an energetic ion track on the
nanoscale where the resulting water flow is cylindrical [7].
Several predictions of this model are very convenient
for benchmarking the simulations performed in the present
work. In particular, the position of the wave front as a
function of time is given by:
R(t) = β√t(cid:20)dT /ds
ρ
(cid:21)1/4
,
(1)
and the pressure of the front as a function of the front
radius is:
Pfront(R) =
1
2 (γ + 1)
β4 dT /ds
R2
.
(2)
In both equations, ρ = 1 g/cm3 is the density of unper-
turbed liquid water, γ = Cp/Cv = 1.222 for liquid water,
dT /ds is the stopping power (i.e., mean energy loss dT
per unit path length ds) of an ion in liquid water, and β
is a parameter which value for liquid water is β = 0.86 [7].
2.2 Molecular dynamics simulation
In the molecular dynamics technique [14] all the atoms of
the system are considered and their classical trajectories
are followed by computing the interaction forces between
all the atoms in the system. The evolution with time of
the coordinates ri(t) of each atom i of mass mi is com-
puted for discrete time steps dt, according to the Langevin
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
3
equation:
mi
d2ri
dt2 =Xj6=i
F ij −
1
τd
mivi + fi ,
(3)
where Pj6=i F ij is the total force acting on atom i as a
consequence of its interaction with all other atoms j in
the system (i.e., Newton's second law). The second and
third terms in the right hand side of Eq. (3) correspond
to the thermostat, used to keep the temperature of the
system nearly constant to T , when coupled to a thermal
bath. In the present work the Langevin thermostat is used,
which exerts a viscous force on each particle of velocity
vi, as well as a random force fi which guarantees thermal
equilibrium. τd is the damping time of the thermostat,
while fi is a Gaussian random force with zero mean and
variance σ2
i = 2mikBT /τd with kB being the Boltzmann's
constant.
For biomolecular systems, where the structure of the
molecule is determined not only by interatomic distances
but also by the geometric configuration of groups of atoms
due to the molecular orbital hybridization, it is common
to use special forcefields describing such interactions. In
the CHARMM forcefield [15], one of the most common
ones for describing biomolecules, the force acting on the
atom i is obtained from the potential energy U (R) as
Pj6=i F ij = dU (R)/dri which corresponds to a given set
of atomic coordinates R and is expressed as a combina-
tion of energies arising from the distances between pairs
of bonded atoms, the angles formed between groups of
three sequentially bonded atoms, the dihedral torsion an-
gle formed by groups of sequentially four bonded atoms,
the improper angles formed between groups of atoms that
should form a plane, and the nonbonded interactions rep-
resented by the pure Coulomb force and the van der Waals
interaction between pairs of atoms:
U (R) = Xbonds
+ Xdihedr.
+Xi Xj6=i
Kb(b − b0)2 + Xangles
Kχ (1 + cos nχ − δ) + Ximprop.
rij (cid:19)12
+"ǫij(cid:18) Rmin,ij
qiqj
ε rij
Kθ(θ − θ0)2 +
(4)
Kϕ(ϕ − ϕ0)2 +
rij (cid:19)6# .
−(cid:18) Rmin,ij
In this equation b is the bond distance between two bonded
atoms, θ is the bond angle between every triplet of sequen-
tially bonded atoms, χ is the dihedral torsion angle formed
by every four atoms connected via covalent bonds and ϕ
is the improper torsion angle, used to maintain planarity
between groups of sequentially bonded atoms; b0, θ0 and
ϕ0 correspond to the equilibrium quantities, while n and δ
determine the periodicity of the dihedral interaction. Kb,
Kθ, Kχ and Kϕ are the corresponding force constants. The
Coulomb interaction is characterized by the atomic partial
charges qi, the interatomic distances rij and the effective
dielectric constant ε. The van der Waals interaction is de-
fined by a 6 -- 12 Lennard-Jones potential with well depth
Table 1. Summary of the different simulation boxes used
for the shock wave simulations. The four systems have a
length of 4.6 nm along the track direction y, where PBC
are applied.
System Shape
Track-to-boundary PBC
distance (nm)
(x, z)
I
II
III
IV
disc
parallelepiped
disc
parallelepiped
10
10
17
17
No
Yes
No
Yes
ǫ and the minimum energy distance Rmin. All of these pa-
rameters can be obtained for many biological molecules,
including nucleic acids and proteins, from the CHARMM
potential [15]. All simulations in this work have been per-
formed using the CHARMM implementation within the
MBN Explorer simulation package [13].
2.2.1 Setting up the biological system
The main constituent of living tissues is liquid water.
Therefore when we refer to the ion-induced shock waves,
they occur mainly in water. However, if we want to es-
timate possible biological effects, we should consider as
a target of the shock wave some biological molecules, for
example DNA. Thus in order to set up a simulation box
for the present study we need to define (i) the water box
in which the shock wave is going to propagate and (ii)
the biological molecule that we are going to use to assess
biodamage.
Point (i), even though it seems straightforward, has to
be considered carefully. This is because the shock wave
is a violent dynamical process and the pressure waves
generated can propagate fast, reaching, under certain cir-
cumstances, the boundaries of the simulation box. Such
"boundary effects" can produce artifacts in the simula-
tions. These artifacts could be simply avoided by setting
a extremely large water box in order to guarantee that the
shock wave will not arrive to the boundaries. However, the
computational cost of molecular dynamics simulations in-
creases with the number of atoms in the simulation box,
so the size of the system has to be selected properly in
order to perform the simulations in reasonable times. For
this reason in the present study we have constructed four
different simulation boxes.
In particular we want to study the effect of the size of
the system in the simulation results as well as the inclusion
of periodic boundary conditions to simulate extended me-
dia. These four systems are summarized in Table 1. The
first box, which we will refer as system I, was designed
as a liquid water disc of 10 nm radius around the track,
located in the disc center. This system is quite convenient
since it is quite small and then computationally efficient.
A radius of 10 nm was chosen following results from the
hydrodynamics model, which shows that the pressure of
the shock wave is already very small after 10 nm [7]. This
disc, 4.6 nm long in the track direction y, was simulated
4
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
with periodic boundary conditions (PBC) only in this di-
rection. In order to check the effect of the inclusion of PBC
in the directions x and z, the parallelepiped system II was
built. It has the same length of 4.6 nm in the track direc-
tion y and the distance from the track to the boundary is
also 10 nm but PBC are applied in the three coordinates.
Finally, to check size effects, two similar systems but with
a track-to-boundary distance of 17 nm where built, sys-
tems III and IV. System IV will be in general used as a
benchmark for the rest of them since it is the largest one
and it has PBC in all directions.
Regarding point (ii), in the present work we would like
to chose a simple system that can be simulated straight-
forwardly for performing more systematic studies of shock
wave effects in biomolecules. As a first approximation we
will use short DNA segments, as it was done in previous
works evaluating the damaging effect of secondary elec-
trons [3]. They are simpler to treat than nucleosomes and,
since shock wave effects are quite local, they should be
large enough for estimating DNA damage.
In the present study we have chosen the B-DNA
molecule from the Protein Data Bank [16] with PDB
ID 309D [17]. This molecule, represented in Fig. 1, con-
sists on a B-DNA duplex containing eight base-pairs and
two free nucleotides at each 5'-end. These sticky ends are
complementary, forming continuous 10-fold double helix
molecules. This molecule is very convenient since it allows
the building of DNA duplex models as long as desired, by
the replication and displacement of the original molecule
as many times as wanted. It opens, as well, the possibility
of being used as building block for more complex DNA
structures if desired.
Currently we have used the 309D molecule to con-
struct a linear DNA duplex, each strand being 30-base
long, which we consider is enough for an initial assess-
ment of the shock wave effects in DNA. After building
this molecule from the original one from the Protein Data
Bank it is solvated in liquid water and an atmosphere of 60
sodium counterions is placed around the DNA (since each
base carries one negative charge) by means of the software
VMD [18]. Then, using MBN Explorer [13], the system is
optimized by a velocity quenching algorithm, and then
equilibrated at T = 310 K (body temperature), using the
Langevin thermostat with damping time τd = 0.2 ps, a
simulation time step dt = 1 fs, periodic boundary con-
ditions, the particle mesh Ewald algorithm for the long-
range Coulomb interactions, and a van der Waals cut off
distance of 13 A. The geometry of the resulting DNA du-
plex, as well as the structure of the sodium ions envi-
ronment, has been checked and compared with reference
data [19,20,21]. The equilibrated DNA duplex has then
been put in the different water boxes described above for
the subsequent shock wave simulations. The DNA duplex,
oriented in the z direction, is always placed with its center
at 2 nm distance from the track, in a way in which the
closest DNA atom is at ∼ 1 nm from the track. The initial
configurations for systems I and II are shown in Figs. 2(a)
and (d).
Fig. 1. (Color online) Representation of the B-DNA
molecule from the Protein Data Bank [16] with PDB ID
309D [17]. Thick lines represent the backbone bonds whose
energy is monitored during the shock wave simulations
(see section 3).
2.2.2 Setting up the initial conditions of the shock wave
An energetic ion losses its energy mainly by electronic ex-
citations and ionizations. As a result a large number of low
energy electrons are produced (below 50 eV in the Bragg
peak) [3,9,10,22] which will propagate on the nanometer
scale, being stopped very quickly. The dynamics of such a
process was studied in Ref. [11], where it was shown how
the radial dose around the ion track is built up in ∼ 50
fs. Also, almost all the energy lost by the ion is deposited
within 1 nm from the ion track. It is well known that
the electron-phonon coupling, the mechanism by which
the energy deposited by the secondary electrons can be
dissipated, occurs in times much longer than fs, i.e., in
the sub-ps scale. This means that a large amount of en-
ergy will be locally deposited within ∼ 1 nm very quickly
(∼ 50 fs) and that it will be released at once, putting the
initial conditions for the formation of the shock wave.
In terms of the molecular dynamics simulations we se-
lect the water molecules initially present within 1 nm ra-
dius from the ion track. These molecules are highlighted
in Fig. 2, where their atoms are shown as spheres. All the
energy lost by the ion (which is not considered explicitly in
the simulations since it crosses the system in much shorter
times) will be transferred to these molecules so the veloci-
ties of their atoms (obtained from previous equilibration)
are multiplied by a factor α, in a way in which their total
kinetic energy after the ion crosses the system is:
1
2
N
Xi
mi(α · vi)2 =
3N kBT
2
· l .
+(cid:12)(cid:12)(cid:12)(cid:12)
dT
ds(cid:12)(cid:12)(cid:12)(cid:12)
(5)
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
5
Fig. 2. (Color online) Snapshots of a shock wave induced by a carbon ion in the Bragg peak region in liquid water with
the track (y direction) crossing systems I and II at 1 nm distance from the DNA duplex (oriented in the z direction).
Panels (a), (b), and (c) correspond to the system I at 0, 5, and 10 ps after the ion traversal, while panels (d), (e), and
(f) represent the same times for the system II. The water molecules initially excited by the ion track as well as the
DNA duplex are highlighted, with their atoms shown as explicit atoms.
The first term in the right hand side of the equation corre-
sponds to the initial kinetic energy of the excited cylinder
(with N atoms) at equilibrium (T = 310 K). The second
term is the energy lost by the ion when crossing the sys-
tem, which is its stopping power, dT /ds times the length
of the simulation box, l. The simulation of the shock wave
is done as indicated previously for the equlibration but
without thermostat and with lower values of the time step
dt, depending on the simulation, to ensure conservation of
energy.
6
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
3 Results and discussion
We will start our study of the effects of ion-induced shock
waves in DNA duplexes with a carbon ion in the Bragg
peak region (with kinetic energy of ∼300 keV/u), carbon
being a common choice in modern ion beam therapy cen-
ters. The stopping power of a carbon ion in the Bragg
peak is 900 eV/nm. Figures 2(a), (b), and (c) show three
snapshots of the evolution of the system after irradiation
(0, 5, and 10 ps, respectively) when using system I, i.e.,
a 10 nm radius disc in vacuum (see Table 1). The same
results are shown in Figs. 2(d), (e), and (f) for system II,
i.e., a parallelepiped of 10 nm track-to-boundary distance
with periodic boundary conditions (PBC).
The explosion of the system is much more violent for
system I than for system II with the structure of the DNA
duplex being heavily distorted in Fig. 2(c). This is due to
the high pressures produced by the shock wave that force
the system to expand into vacuum in system I. However,
this behavior does not occur in system II, apparently be-
cause the periodic boundary conditions provide the pres-
sure to damp the effects of the shock wave. The final struc-
ture of the DNA duplex is distorted in system II (Fig. 2(f))
but not that much as in the case of system I. Therefore,
periodic boundary conditions are needed to suppress the
system explosion if a small system is used for the simula-
tions.
A better understanding of this situation can be achieved
by calculating the pressure generated by the shock wave.
These values can be compared to the results provided by
the analytical model (subsection 2.1). To calculate the
pressure, virtual walls have been placed at different ra-
dial distances r from the track. At several times t during
the simulation the number of atoms crossing this wall in
each direction has been monitored and their momentum
pi = mivi calculated. The pressure is calculated as:
P =
dp
dt A
=
2(cid:16)Pi pi −Pj pj(cid:17)
dt A
,
(6)
where dt is the time passed between frames in the sim-
ulation, A is the surface of the cylindrical wall and the
indexes i and j refer to the atoms crossing the wall in the
outer and in the inner directions respectively. The factor
2 comes from the assumption that, if the wall was there
to measure the pressure, the atoms would have an elas-
tic collision coming back after the collision with the same
momentum pi but in nearly opposite direction, so the mo-
mentum transfer would be dpi ≃ 2pi.
The results for the time evolution of the pressure wave
produced by a carbon ion in the Bragg peak in system
IV (where boundary effects are not expected) are shown
in Fig. 3(a). The wave front can be clearly identified dur-
ing the first picosecond, where the Gaussian shape is very
visible, with a sharp maximum, that propagates rapidly
(faster than ∼1600 m/s) in the radial direction. After the
first few picoseconds the pressure wave widens and loses
intensity quickly. This time evolution can be compared
with the results of the hydrodynamics model (subsection
2.1). This comparison is shown for the position of the front
8
6
4
2
0
25
20
15
10
5
0
)
2
/
m
n
N
n
(
P
(a)
(b)
0.25 ps
0.5 ps
1 ps
2 ps
3 ps
6 ps
9 ps
0
2
4
6
10
12
14
16
8
r (nm)
Fig. 3. (Color online) Time evolution of the pressure, as a
function of the radius from the track r, generated by (a) a
carbon ion and (b) an iron ion in the Bragg peak region in
system IV, a parallelepiped with 17 nm track-to-boundary
distance and periodic boundary conditions (see Table 1).
and its pressure, respectively, in Figs. 4(a) and (b), where
lines are the results from the analytical model and sym-
bols represent the results from the simulations (which have
been obtained by finding the maximum of the Gaussian
fitting each curve in Fig. 3(a)).
The first important observation is that simulation re-
sults in system IV follow the results of the analytical
model for carbon. However, artifacts appear in the re-
sults of systems II and III due to their small size or the
lack of PBC. In the case of system II, where the track-to-
boundary distance is just 10 nm, the tail of the pressure
wave reaches the boundary after ∼3 ps. This produces
a reflection of the pressure wave in the boundary, which
stops the front propagation. Even though the disc-like sys-
tem III has a track-to-boundary distance of 17 nm, the
lack of PBC seems to produce an unnatural propagation
of the front at longer times. Neither of these effects are ob-
served for carbon when system IV is used, what suggests
that the system has to be large enough and PBC should
be applied in order to properly contain the shock wave.
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
7
t (ps)
4
6
8
10
2
0
(a)
10
8
6
4
2
0
25
20
15
10
(b)
C (MD, System II)
C (MD, System III)
C (MD, System IV)
C (Hydrodynamics)
Fe (MD, System IV)
Fe (Hydrodynamics)
t
n
u
o
c
e
v
i
t
l
a
e
R
(a)
100
10-1
10-2
10-3
10-4
10-5
100
(b)
10-1
10-2
10-3
10-4
0.6
0.4
0.2
)
V
e
(
d
n
o
b
U
0.0
0
2
4
6
t (ps)
8
10
C in system I
C in systemII
C in system III
C in system IV
Fe in system IV
C ion in nucleosome [6]
Fe in nucleosome [6]
)
m
n
(
R
)
2
/
m
n
N
n
(
t
n
o
r
f
P
5
0
0
2
4
6
R (nm)
8
10
12
Fig. 4. (Color online) (a) Time evolution of the wave front
position for carbon and iron ion induced shock waves in
the Bragg peak region. (b) Pressure of the wave front as a
function of its radial position, for carbon and iron induced
shock waves in the Bragg peak region. Symbols represent
molecular dynamics results, while lines are the predictions
of the analytical hydrodynamics model. See the text for
further details.
10-5
0
1
2
3
Ubond (eV)
4
5
Fig. 5. (Color online) Relative count of energy deposition
events in the DNA backbone bonds located within 2.2 nm
of the ion track. (a) Present results for a carbon ion in the
Bragg peak region in systems I -- IV. (b) Present results
for carbon and iron ions in system IV, compared with
previous results in the nucleosome [6]. The inset shows
the time evolution of the potential energy of some of the
bonds analyzed for a carbon ion in system IV.
Before discussing results for a heavier ion (iron), we
will analyze the effects of the carbon-induced shock wave
in the DNA duplex. We have monitored the energy stored
in the covalent bonds of the DNA backbone closest to the
track [6]. The inset in Fig. 5(a) shows the potential en-
ergy of some of the covalent bonds (first term in the right
hand side of Eq. (5)), initially located within 2.2 nm from
the track, as a function of time. The energy stored in the
bonds varies in time, with some sudden jumps due to the
exposure of the bonds to the pressure of the shock wave.
Each of those local maxima can be stored as an energy
deposition event. The frequency count of such events for
the shock wave produced by a carbon ion in the Bragg
peak in all the systems I-IV is represented in Fig. 5(a).
This frequency, which presents an exponential behavior,
can be parameterized to estimate the probability for in-
ducing single strand breaks, regarded as energy deposition
larger than some given threshold, typically around 2.5 eV
[6]. Clearly, system I presents a very different behavior
while systems II-IV converge. This result arises from the
artificial violent explosion of system I, due to the lack
of pressure to damp the shock wave. The convergence be-
tween systems II-IV emphasizes the importance of the first
picoseconds of the shock wave in terms of DNA damage:
even though the pressure wave differs in these systems af-
ter ∼ 3 ps, their energy deposition profile is similar, high-
lighting the fact that it is built up during the first picosec-
onds after the explosion. The energy deposition profile is
compared with previous simulations in the nucleosome [6]
in Fig. 5(b), where we find a fairly good match, allowing
for the different geometries.
8
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
From these simulations it is clear that there is a rel-
atively low probability of producing single strand breaks
by the shock wave in the case of carbon ions (assuming a
SSB threshold of 2.5 eV; however, such thresholds can be
even lower [6]). In terms of a possible experimental verifi-
cation of these results it would be convenient to study ions
with larger stopping powers since these can produce larger
numbers of strand breaks, so they are easier to detect. For
this reason we have performed a simulation of the shock
wave produced by an iron ion in the Bragg peak region,
having a stopping power of 7195 eV/nm. System IV has
been used for this simulation simulation since, from the
results for carbon, the other systems would be too small.
The evolution of the pressure wave for iron is shown in
Fig. 3(b), the evolution of the front is depicted in Fig. 4,
and the energy deposition profile in the DNA backbone
bonds is shown in Fig. 5(b). The final geometry of the
system after 10 ps, as compared with the carbon case, is
shown in Fig. 6.
Both the position of the front and its pressure, shown
in Fig. 4, seem to follow the predictions of the analytical
model but only up to 4 ps. After that, it can be clearly seen
that the front actually goes backwards. This behaviour
is caused by the reflection of the pressure wave at the
boundaries of the system, as it happened for carbon with
system II. As expected from Eqs. (1) and (2), the velocity
of propagation of the iron shock wave is 1.682 faster than
for carbon and the pressure of the front is 8 times larger.
This results from the fact that the stopping power for iron
in the Bragg peak region is 8 times larger than for carbon.
Such large pressures lead to larger distortions of the
DNA duplex being observed for iron in comparison with
carbon in Fig. 6. This leads to a larger number of high en-
ergy deposition events in DNA backbone bonds, as shown
in Fig. 5(b), where the slope of the curve is much smaller
than for carbon ion and many more events larger than 2.5
eV (a conservative estimation for the threshold for single
strand break production) are produced. This would justify
the use of heavier ions for a possible experimental verifica-
tion of shock wave effects, where more single strand breaks
would be detected. The results for iron are compared with
previous simulations performed for nucleosome [6]. In this
case, the present results are not that close to the previ-
ously reported values. However, the order of magnitude
seems to be similar. Such differences could be due to sev-
eral reasons, the most probable being the different geom-
etry. The larger slope of the previously reported data sug-
gests that the histone protein has some protective role in
DNA damage. In the present work the DNA duplex is free,
allowing a larger stretching of the molecule which might
be impeded in the nucleosome. Also the fact that system
IV is not large enough for the iron-induced shock wave
could have some contribution. However, as seen from the
results for carbon, the first picoseconds of the simulation
are the most relevant for the DNA damage and the molec-
ular dynamics simulation follows properly the analytical
results up to 4 ps. It is clear, in any case, that a larger
system is needed for the simulation of an iron ion shock
wave. This has not been done here in order not to increase
Fig. 6. (Color online) Geometry of system IV, 10 ps af-
ter passage of (a) a carbon ion and (b) an iron ion, in
the Bragg peak region. Nanochannels are formed in liquid
water, and the DNA structure is distorted, particularly
heavily in the case of the iron ion.
the computational expense of the calculations. However,
precautions should be taken in future calculations to op-
timize the box for each ion. In any case, the main conclu-
sion that can be extracted from Fig. 4 is that molecular
dynamics simulations can reproduce perfectly the results
predicted by the analytical hydrodynamics model as long
as the system used for the simulations is large enough for
containing the wave and PBC are applied. Indeed the hy-
drodynamics model can be used as a consistency check to
analyze whether the system built for molecular dynamics
simulation is appropriate or not.
4 Conclusions
In this paper we have presented a molecular dynamics
study of the effects of ion-induced shock waves in biologi-
cal media. The study focuses on the pressure waves arising
from the heating of liquid water after energetic carbon and
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
9
iron ions traversal (in the Bragg peak region) and on the
effects of such waves on short DNA duplexes.
The first important conclusion of this paper is that
molecular dynamics simulations of
ion-induced shock
waves can reproduce the results obtained by the analyti-
cal hydrodynamics model previously reported, in terms of
front velocity and pressure, as long as the simulation box
is properly designed. This point is of great importance:
ion-induced shock waves have not been detected yet ex-
perimentally and their existence has been so far only pre-
dicted theoretically. The fact that independent techniques,
such as classical molecular dynamics and classical hydro-
dynamics, coincide in the prediction of the properties of
the shock waves adds arguments in favor of their existence
and establishes the theoretical grounds over which possi-
ble experiments can be designed and interpreted for their
confirmation.
Molecular dynamics simulations reproduce the analyt-
ical results only if the system is well designed. This means
that it has to be large enough to contain the pressure
waves produced and that preferably periodic boundary
conditions should be applied to avoid boundary effects
in the simulation, such as the reflection of the pressure
wave in the periodic boundaries or the expansion of the
system into vacuum when PBC are not applied. In the
present case, a track-to-boundary distance of 17 nm has
been demonstrated to be large enough for carbon-induced
shock waves in the Bragg peak region. For iron larger
boxes should be used in future work. However, smaller
boxes can be used for certain simulations as long as one
is interested in short times and distances from the track
where most of the damage occurs. The analytical hydro-
dynamics model can be always used to check the consis-
tency of the results, as shown in this paper. This is some-
thing that has not been considered before and has not
been checked in other works.
The energy deposition profile in the DNA backbone
bonds has been analyzed, both for carbon and iron ions,
as an estimation of possible biological damage. The results
obtained for carbon are very similar to those reported pre-
viously for molecular dynamics simulations in the nucle-
osome [6]. For the case of iron ions we have found some
differences with the previous published data. However, the
order of magnitude of the results is quite similar, confirm-
ing the systematics of the stopping power of the ion on
the expected damage of DNA. The observed differences
are most likely due to the different geometry of the sys-
tems used, where the presence of the histone protein in
the nucleosome can have some protective effect.
The present results establish a solid procedure for per-
forming more systematic simulations of shock wave effects
in DNA duplexes, where the analytical hydrodynamics
model can be used as a benchmark. The properties of
the shock waves for ions of different stopping powers can
be predicted and their effects on DNA determined. This
opens the door to new simulation improvements, such as
the inclusion of reactive force fields for a better prediction
of DNA damage [23,24] or the study of the shock wave
effects in the propagation of secondary species generated
around ion tracks. Such systematic studies will allow a
better understanding of the biological relevance of ion-
induced shock waves and will be useful for the future de-
sign and interpretation of potential experiments for their
detection.
The authors acknowledge financial support from the European
Unions FP7-People Program (Marie Curie Actions) within the
Initial Training Network No. 608163 "ARGENT", Advanced
Radiotherapy, Generated by Exploiting Nanoprocesses and
Technologies. The molecular dynamics simulations were per-
formed at the computer clusters IMACT, at The Open Univer-
sity (UK), and KELVIN, at Queen's University Belfast (UK).
References
1. J. S. Loeffler and M. Durante, Nature Rev. Clin. Oncol. 10,
411 (2013)
2. M. Scholz and T. Elsasser, Adv. Space Res. 40, 1381 (2007)
3. E. Surdutovich and A. V. Solov'yov, Eur. Phys. J. D 68,
353 (2014)
4. D. Schardt, T. Elsasser, and D. Schulz-Ertner, Rev. Mod.
Phys. 82, 383 (2010)
5. B. Boudaıffa, P. Cloutier, D. Hunting, M. A. Huels, and L.
Sanche, Science 287, 1658 (2000)
6. E. Surdutovich, A. V. Yakubovich, and A. V. Solov'yov,
Scientific Reports 3, 1289 (2013)
7. E. Surdutovich and A. V. Solov'yov, Phys. Rev. E 82,
051915 (2010)
8. A. V. Yakubovich, E. Surdutovich, and A. V. Solov'yov,
Nucl. Instr. Meth. B 297, 135 (2012)
9. P. de Vera, R. Garcia-Molina, I. Abril, and A. V. Solov'yov,
Phys. Rev. Lett. 110, 148104 (2013)
10. P. de Vera, I. Abril, R. Garcia-Molina, and A. V. Solov'yov,
J. Phys.: Conf. Ser. 438, 012015 (2013)
11. E. Surdutovich and A. V. Solov'yov, Europ. Phys. J. D 69,
193 (2015)
12. L. Landau and E. Lifshitz, Fluid Dynamics, 2nd ed. (Reed-
Elsevier, Oxford, Boston, Johannesburg, 1987), Vol. 6.
13. I. A. Solov'yov, A. V. Yakubovich, P. V. Nikolaev, I.
Volkovets, and A. V. Solov'yov, J. Comp. Chem. 33, 2412
(2012); www.mbnexplorer.com
14. M. P. Allen and D. J. Tildesley, Computer simulation of
liquids (Oxford University Press, Oxford, 1989).
15. A. D. MacKerell, Jr., D. Bashford, M. Bellott, R. L. Dun-
brack, Jr., J. D. Evanseck, et al., J. Phys. Chem. B 102, 3586
(1998)
16. H. M. Berman, J. Westbrook, Z. Feng, G. Gilliland, T. N.
Bhat, H. Weissig, I. N. Shindyalov, P. E. Bourne, Nucl. Acids
Res. 28, 235 (2000)
17. H. Qiu,
Biol.
J. C. Dewan, N. C.
267,
881
Mol.
http://www.rcsb.org/pdb/explore/explore.do?structureId=309D.
(1997). Available
Seeman,
J.
online:
18. W. Humphrey, A. Dalke, and K. Schulten, J. Molec.
Graphics 14, 33 (1996)
19. U. Heinemann, C. Alings and M. Bansal, The EMBO Jour-
nal 11, 1931 (1992)
20. F. Pan, C. Roland and C. Sagui, Nucl. Acids Res. 42,
13981 (2004)
21. T. J. Robbins and Y. Wang, J. Biomolec. Struct. Dynam.
31, 1311 (2013)
10
P. de Vera et al.: Molecular dynamics study of ion-induced shock waves in biological media
22. O.I. Obolensky, E. Surdutovich, I. Pshenichnov, I. Mishus-
tin, A.V. Solov'yov, and W. Greiner, Nucl. Instr. Meth. B
266, 1623 (2008)
23. D. Bottlander, C. Muckschb, H. M. Urbassek, Nucl. Instr.
Meth. B 365, 622 (2015)
24. G. B. Sushko, I. A. Solov'yov, A. V. Verkhovtsev, S. N.
Volkov, and A. V. Solov'yov, Eur. Phys. J. D 70, 12 (2016).
|
1910.05413 | 1 | 1910 | 2019-10-11T21:06:12 | Temporal precision of molecular events with regulation and feedback | [
"physics.bio-ph",
"q-bio.MN"
] | Cellular behaviors such as migration, division, and differentiation rely on precise timing, and yet the molecular events that govern these behaviors are highly stochastic. We investigate regulatory strategies that decrease the timing noise of molecular events. Autoregulatory feedback increases noise. Yet, we find that in the presence of regulation by a second species, autoregulatory feedback decreases noise. To explain this finding, we develop a method to calculate the optimal regulation function that minimizes the timing noise. The method reveals that the combination of feedback and regulation minimizes noise by maximizing the number of molecular events that must happen in sequence before a threshold is crossed. We compute the optimal timing precision for all two-node networks with regulation and feedback, derive a generic lower bound on timing noise, and discuss our results in the context of neuroblast migration during Caenorhabditis elegans development. | physics.bio-ph | physics | Temporal precision of molecular events with regulation and feedback
Shivam Gupta,1 Sean Fancher,1, 2 Hendrik C. Korswagen,3 and Andrew Mugler1, ∗
1Department of Physics and Astronomy, Purdue University, West Lafayette, IN 47907, USA
2Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104, USA
3Hubrecht Institute, Royal Netherlands Academy of Arts and Sciences
and University Medical Center Utrecht, 3584 CT Utrecht, Netherlands
Cellular behaviors such as migration, division, and differentiation rely on precise timing, and yet
the molecular events that govern these behaviors are highly stochastic. We investigate regulatory
strategies that decrease the timing noise of molecular events. Autoregulatory feedback increases
noise. Yet, we find that in the presence of regulation by a second species, autoregulatory feedback
decreases noise. To explain this finding, we develop a method to calculate the optimal regulation
function that minimizes the timing noise. The method reveals that the combination of feedback
and regulation minimizes noise by maximizing the number of molecular events that must happen in
sequence before a threshold is crossed. We compute the optimal timing precision for all two-node
networks with regulation and feedback, derive a generic lower bound on timing noise, and discuss
our results in the context of neuroblast migration during Caenorhabditis elegans development.
Precise timing is crucial for many biological processes
including cell division [1 -- 3], cell differentiation [4], cell
migration [5], embryonic development [6, 7], and cell
death [8]. Ultimately the timing of these processes is gov-
erned by the timing of molecular events inside the cell.
However, these events are inherently stochastic. Cells use
regulatory networks to reduce this stochasticity, but the
effects of particular regulatory features on timing pre-
cision remain poorly understood. We recently demon-
strated that the time at which an accumulating molecu-
lar species crosses an abundance threshold is more precise
if that species is regulated by a second species with its
own stochastic dynamics [9]. In contrast, it was recently
demonstrated that if the species is instead regulated by
itself (feedback), then the crossing time is less precise
[10]. Yet, feedback is common in many important tim-
ing processes.
In yeast, the cyclin proteins that cross
an abundance threshold to initiate the cell cycle [3] are
subject to positive feedback [1, 11, 12]. In Caenorhabdi-
tis elegans, the mig-1 protein that crosses an abundance
threshold to terminate migration in QR neuroblasts [5]
has been found in experiments on the sister QL lineage
to be subject to feedback via Wnt signaling [13]. This
raises the question of why feedback is observed in key
timing processes if it has been shown to decrease timing
precision.
Here we investigate the combined effect of regulation
and feedback on timing precision. We develop a gradient-
descent approach to find the globally optimal regulation
function for a given network topology that minimizes
the timing noise. We find that, despite the fact that
feedback generically increases timing noise when it acts
alone, feedback decreases timing noise when it acts in
combination with regulation by an external species. We
explain the mechanisms behind this counterintuitive re-
sult, derive a generic lower bound on the timing noise,
and discuss the relevance of our results to the timing of
neuroblast migration in C. elegans.
FIG. 1: Feedback increases timing precision in the presence
but not absence of regulation.
(A) A species Y crosses a
molecule-number threshold y∗ at mean time t∗ with timing
variance σ2
t . (B) Feedback increases the variance. However, in
the presence of regulation by a second species X, feedback on
either (C) Y or (D) X can decrease the variance. Parameters
are Ky = 2.5 in B; α0t∗ = 10, Hx = −0.5, Hxy = −Hy,
Kx = 15, Ky = 5, and Kxy = 6 in C; α0t∗ = 10, Hy = 4,
Kx = 10, and Ky = 7.5 in D; and y∗ = 10 throughout.
Consider a molecular species Y that is produced over
time and first reaches a molecule-number threshold y∗
at a particular time t∗ on average (Fig. 1A). Stochastic-
ity in the accumulation process leads to variability in the
crossing time t. The timing noise is given by the variance
σ2
t . For unregulated production of Y , the time between
each production event is exponentially distributed with
9
1
0
2
t
c
O
1
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
3
1
4
5
0
.
0
1
9
1
:
v
i
X
r
a
012051015-1010.511.50240.50.60.70.80.91-8-6-4-200.50.60.70.80.91ABCDYXHy<latexit sha1_base64="iurlgOJ3OkaVVILAwpRNzkWzs0c=">AAAB6nicbVBNS8NAEJ34WetX1aOXxSJ4KkkV9Fj00mNF+wFtKJvtpF262YTdjRBKf4IXD4p49Rd589+4bXPQ1gcDj/dmmJkXJIJr47rfztr6xubWdmGnuLu3f3BYOjpu6ThVDJssFrHqBFSj4BKbhhuBnUQhjQKB7WB8N/PbT6g0j+WjyRL0IzqUPOSMGis91PtZv1R2K+4cZJV4OSlDjka/9NUbxCyNUBomqNZdz02MP6HKcCZwWuylGhPKxnSIXUsljVD7k/mpU3JulQEJY2VLGjJXf09MaKR1FgW2M6JmpJe9mfif101NeONPuExSg5ItFoWpICYms7/JgCtkRmSWUKa4vZWwEVWUGZtO0YbgLb+8SlrVindZqd5flWu3eRwFOIUzuAAPrqEGdWhAExgM4Rle4c0Rzovz7nwsWtecfOYE/sD5/AEw/o28</latexit>XYHx<latexit sha1_base64="LpYWhs7E/rE/PxiN3jaT+6oN590=">AAAB6nicbVDLTgJBEOzFF+IL9ehlIjHxRHbRRI9ELxwxyiOBDZkdemHC7OxmZtZICJ/gxYPGePWLvPk3DrAHBSvppFLVne6uIBFcG9f9dnJr6xubW/ntws7u3v5B8fCoqeNUMWywWMSqHVCNgktsGG4EthOFNAoEtoLR7cxvPaLSPJYPZpygH9GB5CFn1FjpvtZ76hVLbtmdg6wSLyMlyFDvFb+6/ZilEUrDBNW647mJ8SdUGc4ETgvdVGNC2YgOsGOppBFqfzI/dUrOrNInYaxsSUPm6u+JCY20HkeB7YyoGeplbyb+53VSE177Ey6T1KBki0VhKoiJyexv0ucKmRFjSyhT3N5K2JAqyoxNp2BD8JZfXiXNStm7KFfuLkvVmyyOPJzAKZyDB1dQhRrUoQEMBvAMr/DmCOfFeXc+Fq05J5s5hj9wPn8AL3qNuw==</latexit>YHy<latexit sha1_base64="iurlgOJ3OkaVVILAwpRNzkWzs0c=">AAAB6nicbVBNS8NAEJ34WetX1aOXxSJ4KkkV9Fj00mNF+wFtKJvtpF262YTdjRBKf4IXD4p49Rd589+4bXPQ1gcDj/dmmJkXJIJr47rfztr6xubWdmGnuLu3f3BYOjpu6ThVDJssFrHqBFSj4BKbhhuBnUQhjQKB7WB8N/PbT6g0j+WjyRL0IzqUPOSMGis91PtZv1R2K+4cZJV4OSlDjka/9NUbxCyNUBomqNZdz02MP6HKcCZwWuylGhPKxnSIXUsljVD7k/mpU3JulQEJY2VLGjJXf09MaKR1FgW2M6JmpJe9mfif101NeONPuExSg5ItFoWpICYms7/JgCtkRmSWUKa4vZWwEVWUGZtO0YbgLb+8SlrVindZqd5flWu3eRwFOIUzuAAPrqEGdWhAExgM4Rle4c0Rzovz7nwsWtecfOYE/sD5/AEw/o28</latexit>mean t∗/y∗ and variance (t∗/y∗)2. Because the produc-
tion events are independent, the variances add, giving a
t = y∗(t∗/y∗)2 = t2∗/y∗. Therefore we
total variance of σ2
t y∗/t2∗, whose value is 1 for
focus on the scaled variance σ2
unregulated production.
First we investigate the effect of feedback on timing
precision using a simple example: we suppose that the
production rate of Y is not a constant but rather is a sim-
ple sigmoidal function of the current number of molecules
y,
β(y) = β0{1 + tanh[Hy(y/Ky − 1)]},
(1)
where positive (negative) Hy corresponds to positive
(negative) feedback, Hy is the maximum steepness, Ky
is the molecule number at which β is half-maximal, and
β0 is set to ensure that the average time at which y first
reaches y∗ is t∗. We calculate the variance σ2
t from the
master equation by matrix inversion [9]. In Fig. 1B we
see that when there is no feedback (Hy = 0), the vari-
t y∗/t2∗ = 1, and that either positive or
ance satisfies σ2
negative feedback increases the variance. This result is
consistent with previous findings for a species that does
not degrade [10], and it has an intuitive explanation: a
sequence of time-ordered stochastic events is most pre-
cisely timed if the mean time for each event to occur is
equal, but feedback makes these times unequal.
Next we investigate the interplay of feedback and regu-
lation by introducing a second species X that is produced
at a constant rate α0. The Y production rate β(x, y) is
now a function of both molecule numbers x and y. We
find that if it is a simple sum β(x, y) = f1(x) + f2(y) or
product β(x, y) = f1(x)f2(y) then feedback continues to
generically increase the timing variance, but if we include
a coupling term β(x, y) = f1(x)f2(y)f3(xy) the situation
is different. Specifically, Fig. 1C shows the case where
β(x, y) = β0{1 + tanh[Hx(x/Kx − 1)]}
× {1 + tanh[Hy(y/Ky − 1)]}
× {1 + tanh[Hxy(xy/K 2
xy − 1)]}.
(2)
We see that with no feedback (Hy = 0) we have
t y∗/t2∗ < 1, which demonstrates that regulation by a
σ2
second species increases the timing precision as found
previously [9]. However, now we also see that with pos-
itive feedback (Hy > 0), the variance can be even lower.
Together with Fig. 1B, this result implies that although
feedback increases timing noise in the absence of regu-
lation, it can decrease timing noise in the presence of
regulation.
Similarly we investigate the case where the feedback
occurs on X, not Y . We take the production rates of x
and y to be
α(x) = α0{1 + tanh[Hx(x/Kx − 1)]},
β(x) = β0{1 + tanh[Hy(x/Ky − 1)]},
(3)
(4)
2
respectively. We see in Fig. 1D that with negative feed-
back (Hx < 0) the variance is lower than with no feed-
back (Hx = 0), again implying that feedback can reduce
timing noise when coupled to regulation.
To understand this effect, we develop a gradient-
descent method to find the optimal regulation that min-
imizes the timing variance. The regulation is specified
by the X and Y production rates α(x, y) and β(x, y),
respectively, which each depend on the molecule num-
bers x and y in general, but whose dependencies will
later be restricted to consider particular feedback topolo-
gies. The probability of first reaching y = y∗ at time t is
P (t) =(cid:80){(cid:126)s} P (t(cid:126)s)P ((cid:126)s), where
S−1(cid:89)
(cid:32)S−1(cid:89)
i=0
ri
ki
,
(cid:90) ∞
i=0
0
P ((cid:126)s) =
P (t(cid:126)s) =
(cid:33)
δ
t − S−1(cid:88)
j=0
.
tj
(5)
(6)
−kiti
dtikie
In Eq. 5, P ((cid:126)s) is the probability of taking a path (cid:126)s from
(x0, y0) = (0, 0) to (xS, yS) = (xS, y∗) for any non-
negative xS, where S is the length of the path. Each
step i takes the system out of state (xi, yi) with rate
ki = α(xi, yi) + β(xi, yi) and into a new state with prob-
ability ri/ki, where the new state is either (xi + 1, yi)
with ri = α(xi, yi) or (xi, yi + 1) with ri = β(xi, yi). In
Eq. 6, P (t(cid:126)s) is the probability that traversing the given
path (cid:126)s takes a time t. The first term integrates over all
values of each step's transition time ti, which is expo-
nentially distributed with rate ki, and the second term
ensures that the sum of these transition times is t. From
P (t) we calculate the moments [14], of which the first two
are
(cid:88)
(cid:88)
{(cid:126)s}
{(cid:126)s}
,
1
ki
S−1(cid:88)
(cid:32)S−1(cid:88)
i=0
i=0
P ((cid:126)s)
P ((cid:126)s)
(cid:104)t(cid:105) =
(cid:104)t2(cid:105) =
(cid:33)
+
S−1(cid:88)
j=0
1
kj
2 .
1
k2
i
(7)
(8)
The optimal regulation function minimizes (cid:104)t2(cid:105) at fixed
(cid:104)t(cid:105) = t∗. Therefore, defining a vector (cid:126)γ whose compo-
nents are all components of both the α(x, y) and β(x, y)
matrices, we initialize (cid:126)γ to satisfy (cid:104)t(cid:105) = t∗ and update it
as
(cid:126)γ(n+1) = (cid:126)γ(n) − (cid:126)u.
(9)
Here (cid:28) 1, and (cid:126)u is such that (cid:126)u · ∇γ(cid:104)t2(cid:105) is maximized
with respect to the constraints (cid:126)u·∇γ(cid:104)t(cid:105) = 0 and u2 = 1.
First we apply this method to the case where X reg-
ulates Y with no feedback. Thus, we fix α = α0 and
optimize β(x). Figure 2A shows the result, and we see
that the optimal β(x) is an increasing function of x (i.e.,
3
in Fig. 1C. The first property ensures that Y is not pre-
maturely activated at early times when x is small. The
second property provides an additional acceleration of y
at late times when x is large. Thus, X acts as a "timer"
for Y , allowing Y to apply self-amplification only at late
times. This has two advantages, as seen in Fig. 2D: (i)
it increases the slope of ¯y at crossing, beyond that with-
out feedback; and (ii) it allows the acceleration to begin
at a ¯y value that is already close to y∗, thus reducing
trajectory-to-trajectory variability caused by prolonged
self-amplification [10].
Finally we consider the case where feedback acts on X
instead of Y . Here, to provide a reasonable constraint
on x(t), we introduce a bound x∗ and restrict α(x) such
that ¯x(t) ≤ x∗ over the range 0 ≤ t ≤ t∗. The optimal
regulation functions α(x) and β(x) are shown in Fig. 2E.
We see that X represses itself and activates Y , and that
both regulation functions have a sharp transition when
x = x∗. We see in Fig. 2F that the resulting dynamics
are sharply kinked.
To understand the sharp nature of the optimal solution
in Fig. 2E and F, we investigate our optimization scheme
(Eqs. 5-9) analytically. The analytic version of Eq. 9 is
0 = γi∂γi ((cid:104)t2(cid:105) − λ(cid:104)t(cid:105)), where the Lagrange multiplier λ
enforces (cid:104)t(cid:105) = t∗, and the factor of γi in front enforces
γi > 0 [14]. By inserting Eqs. 7 and 8 into this condition,
we show [14] that it is satisfied when (i) α and β are
such that all possible paths (cid:126)s to reach y = y∗ have the
same length S, and (ii) all transition rates along each
of these paths are equal. Each such set of equal-length,
constant-velocity paths is a local optimum, and the global
optimum that minimizes the timing variance is the set for
which (iii) the path length S is as large as possible. More
generally, if only property (ii) is satisfied, we show [14]
that the timing variance satisfies
σ2
t
t2∗
=
σ2
S(cid:104)S(cid:105)2 +
1
(cid:104)S(cid:105) ,
(10)
where (cid:104)S(cid:105) and σ2
S are the mean and variance of the path
lengths, weighted by the path probabilities P ((cid:126)s). Clearly
S = 0 and (cid:104)S(cid:105) is as large
the variance is minimized when σ2
as possible, consistent with properties (i) and (iii) above,
respectively.
Now we can understand why the the optimal solution
in Fig. 2E and F looks the way it does. The sharp nature
of the regulation functions ensures that at early times
only x changes, and at late times only y changes, con-
fining the stochastic dynamics to only one possible path
in (x, y) space [property (i)]. The values of α and β,
when they are nonzero, are constant and equal to each
other, ensuring that the velocity along this path is con-
stant [property (ii)]. Finally, both x and y attain their
maximal values x∗ and y∗, ensuring that the path is as
long as possible [property (iii)].
Indeed, Fig. 3 shows the optimal solutions for all
of the networks considered thus far in terms of these
FIG. 2: Optimal regulation functions that minimize timing
variance. (A) Without feedback, X activates Y , (B) allow-
ing ¯y to accelerate before crossing y∗. (C) With feedback on
Y , X acts as a "timer" for Y , allowing Y to self-repress at
early times and self-activate at late times, and (D) providing
further, late-phase acceleration of ¯y. (E) With feedback on
X, it represses itself and activates Y sharply, (F) resulting in
kinked dynamics where ¯x and ¯y growth are separated in time.
Parameters are α0t∗ = 7 (A-D), x∗ = 7 (E, F), and y∗ = 10
throughout.
X activates Y ). The reason, clear from the mean dynam-
ics in 2B, is that as x increases over time, β(x) increases
over time, which causes y to accelerate. The accelera-
tion allows ¯y to cross y∗ with a large slope, reducing the
uncertainty of the crossing time. We observed this ef-
fect previously with Hill-function activation [9], but the
optimal regulation function was unknown.
Next we keep α = α0, but we allow feedback on Y and
find the optimal β(x, y). Figure 2C shows the result, and
we see that the optimal β(x, y) depends on y, confirming
that feedback is beneficial in the presence of regulation.
Specifically, we see that β(x, y) decreases with y (negative
feedback) when x is small, and increases with y (positive
feedback) when x is large. These two properties are also
exhibited by Eq. 2 with Hx < 0, Hy > 0, and Hxy < 0 as
ABDCEFXYYXXYYXXY4
(A) Ranking of timing variance for all one- and two-
FIG. 4:
node networks. Global minimum is σ2
t y∗/t2∗ = y∗/(x∗ + y∗).
In red network, link 1, 2, or both is required. Parameters as
in Fig. 2. (B) Mean dynamics and regulation function (inset)
for case when X is degraded. Here α0t∗ = 3.5.
ulated (cyan, orange), this corresponds to replacing its
production propensity α0 (for x → x + 1) with a degra-
dation propensity α0x (for x → x − 1). The resulting
minimal values of the timing variance are shown by the
open circles in Fig. 4A, and we see that they are lower
than the corresponding values when X is produced over
time (filled circles). The reason, illustrated for the case
where X regulates Y in Fig. 4B, is that when X is pro-
duced over time it increases linearly (Fig. 2B dashed),
whereas when X is degraded over time it decreases expo-
nentially (Fig. 4B dashed). The curvature of the expo-
nential begins to approximate the kinked dynamics of the
globally optimal solution (Fig. 2F dashed). Specifically,
X is most dynamic at early times (Fig. 4B dashed), and
Y is only produced once x drops below a particular value
(Fig. 4B inset) allowing it to be most dynamic at late
times (Fig. 4B solid). Thus, even without feedback, the
nonlinear dynamics of a degraded regulator allow its tar-
get to more closely approach the globally optimal timing
precision.
How can these results be tested in experimental sys-
tems? Our findings suggest that a cellular process where
timing precision is important should be governed by a
molecular network with both multistep regulation and
feedback, particularly one in which every species is sub-
ject to regulation as in Fig. 4A (red). An experimen-
tal example in which timing precision is particularly well
studied is neuroblast migration in developing C. elegans
larvae. Here, the QR neuroblast produces a protein
called mig-1 that crosses an abundance threshold to ter-
minate migration; overproduction causes undermigration
and vice versa [5]. It was recently discovered in the sis-
ter QL lineage that mig-1 is subject to both regulation
and negative feedback via canonical Wnt signaling [13].
Specifically, mig-1 activates one or more Wnt signaling
factors, which in turn repress mig-1. These interactions
form a network of the red type in Fig. 4A (with link 2),
where X is the Wnt factor and Y is mig-1, which is pre-
cisely the class of networks that we predict achieve the
FIG. 3: Properties that minimize timing variance: (A) large
path length S, (B) constant velocity v(t) along path, and (C)
small path length variance σ2
S. Parameters as in Fig. 2.
three properties. Specifically, Fig. 3A shows the mean
dynamics in (x, y) space; Fig. 3B shows the velocity
v(t) = (cid:112)(d¯x/dt)2 + (d¯y/dt)2 along this path, normal-
ized by its time average ¯v = t−1∗ (cid:82) t∗
0 dt v(t); and Fig. 3C
shows the variance σ2
S in the path length across all paths.
With only Y and no X (blue), there is only one possible
path (Fig. 3A), and therefore σ2
S = 0 (Fig. 3C). The op-
timal solution has constant velocity along the path (Fig.
3B), which is achieved with no feedback. When X regu-
lates Y (cyan, orange), the mean path extends into the
(x, y) plane (Fig. 3A), which increases its length and
thus lowers the timing variance. However, it also makes
the velocity non-constant (Fig. 3B) and allows for many
possible paths such that σ2
S > 0 (Fig. 3C). Only upon
allowing X to also regulate itself (red) does the path
become as long as possible (Fig. 3A), constant-velocity
(Fig. 3B), and unique (Fig. 3C).
The minimal values of the timing variance for the net-
works are shown by the filled circles in Fig. 4A. We see
t y∗/t2∗ =
that the single species Y achieves the standard σ2
1 (blue), regulation by X lowers the variance (cyan), feed-
back on Y lowers it further (orange), and regulation of
X lowers it to the global minimum given by Eq. 10 with
S = 0 and (cid:104)S(cid:105) = x∗+y∗, namely σ2
t y∗/t2∗ = y∗/(x∗+y∗).
σ2
Because the results in Fig. 4A are minima, it does not
matter in the last case whether the regulation of X is
by X itself (red link 1), by Y (red link 2), or both; the
optimal regulation functions will produce the red path in
Fig. 3 regardless.
Thus far we have only considered the scenario where X
is produced over time. However, X could alternatively be
degraded over time [9]. In the cases where X is unreg-
0123456780123456789101100.510.811.21.40246ABCYXYYXXYYXYYXXY00.51024681005010200.50.60.70.80.91YXYYXXY21ABXYglobally minimum timing noise. We anticipate that other
biological processes where timing precision is paramount
will be governed by interaction networks in this class.
We have developed a gradient-descent approach that
provides the optimal regulation functions for a given
network topology that minimize the timing noise of a
threshold-crossing event. The approach has revealed that
feedback reduces timing noise in the presence but not ab-
sence of regulation because the combination of the two
increases the number of transitions that must happen
sequentially in molecular state space. More generally,
our work suggests a perspective where noise is not mini-
mized by finding the right network topology, but rather
by finding the right combination of regulation functions
that produce a path through state space that is as long,
steady, and unique as possible. Our approach is straight-
forward to generalize to larger and more complex net-
works, and we anticipate that this perspective applies
broadly to biological processes where timing is crucial.
This work was supported by Human Frontier Science
Program grant RGP0030/2016 and Simons Foundation
grant 376198.
∗ Electronic address: [email protected]
5
[1] J. M. Bean, E. D. Siggia, and F. R. Cross, Molecular cell
21, 3 (2006).
[2] I. Nachman, A. Regev, and S. Ramanathan, Cell 131,
544 (2007).
[3] B. L. Schneider, J. Zhang, J. Markwardt, G. Tokiwa,
T. Volpe, S. Honey, and B. Futcher, Molecular and cel-
lular biology 24, 10802 (2004).
[4] K. Carniol, P. Eichenberger, and R. Losick, Journal of
Biological Chemistry 279, 14860 (2004).
[5] R. A. Mentink, T. C. Middelkoop, L. Rella, N. Ji, C. Y.
Tang, M. C. Betist, A. van Oudenaarden, and H. C. Ko-
rswagen, Developmental cell 31, 188 (2014).
[6] H. Meinhardt, Models of biological pattern formation
(Academic Press Inc, 1982).
[7] D. E. Tufcea and P. Fran¸cois, Biophysical journal 109,
1724 (2015).
[8] J. Roux, M. Hafner, S. Bandara, J. J. Sims, H. Hudson,
D. Chai, and P. K. Sorger, Molecular systems biology 11,
803 (2015).
[9] S. Gupta, J. Varennes, H. C. Korswagen, and A. Mugler,
PLoS computational biology 14, e1006201 (2018).
[10] K. R. Ghusinga, J. J. Dennehy, and A. Singh, Pro-
ceedings of the National Academy of Sciences 114, 693
(2017).
[11] L. Dirick and K. Nasmyth, Nature 351, 754 (1991).
[12] F. R. Cross and A. H. Tinkelenberg, Cell 65, 875 (1991).
[13] N. Ji, T. C. Middelkoop, R. A. Mentink, M. C. Betist,
S. Tonegawa, D. Mooijman, H. C. Korswagen, and A. van
Oudenaarden, Cell 155, 869 (2013).
[14] See Supplementary Material.
SUPPLEMENTARY MATERIAL
Calculation of the moments of the first passage time
Using Eqs. 5 and 6 of the main text, we write the first passage time distribution as
{(cid:126)s}
(cid:88)
(cid:88)
(cid:88)
{(cid:126)s}
P ((cid:126)s)P (t(cid:126)s)
(cid:32)S−1(cid:89)
(cid:32)S−1(cid:89)
i=0
(cid:33)S−1(cid:89)
(cid:90) ∞
ri
ki
j=0
dtirie
(cid:90) ∞
0
−kj tj
δ
t − S−1(cid:88)
tj
(cid:33)
t(cid:96)
(cid:32)
t − S−1(cid:88)
.
(cid:96)=0
dtjkje
(cid:33)
−kiti
δ
j=0
{(cid:126)s}
i=0
0
0
dt tnP (t)
(cid:90) ∞
(cid:32)S−1(cid:89)
(cid:90) ∞
dt tn(cid:88)
(cid:32)S−1(cid:89)
(cid:90) ∞
(cid:88)
{(cid:126)s}
i=0
0
{(cid:126)s}
i=0
0
(cid:90) ∞
0
−kiti
dtirie
(cid:33)
n
δ
t − S−1(cid:88)
j=0
tj
.
−kiti
dtirie
(cid:33)S−1(cid:88)
j=0
tj
The nth moment is
P (t) =
=
=
(cid:104)tn(cid:105) =
=
=
(11)
(12)
Specifically, the first and second moments are
(cid:33)S−1(cid:88)
j=0
tj
−kiti
dtirie
i=0
(cid:32)S−1(cid:89)
(cid:90) ∞
(cid:33) S−1(cid:88)
(cid:32)S−1(cid:89)
S−1(cid:88)
ri
ki
j=0
i=0
0
P ((cid:126)s)
1
kj
j=0
{(cid:126)s}
(cid:88)
(cid:88)
(cid:88)
{(cid:126)s}
{(cid:126)s}
1
kj
(cid:104)t(cid:105) =
=
=
and
(cid:10)t2(cid:11) =
=
=
=
=
{(cid:126)s}
{(cid:126)s}
(cid:88)
(cid:88)
(cid:88)
(cid:88)
(cid:88)
{(cid:126)s}
{(cid:126)s}
{(cid:126)s}
−kiti
−kiti
0
0
i=0
i=0
dtirie
dtirie
(cid:32)S−1(cid:89)
(cid:32)S−1(cid:89)
(cid:32)S−1(cid:89)
(cid:32)S−1(cid:89)
(cid:90) ∞
(cid:90) ∞
(cid:33) S−1(cid:88)
S−1(cid:88)
(cid:33)
S−1(cid:88)
S−1(cid:88)
+
P ((cid:126)s)
ri
ki
ri
ki
j=0
j=0
i=0
i=0
(cid:96)=j
1
k2
j
1
k2
j
j=0
tj
j=0
j=0
j=0
t2
j +
(cid:33)S−1(cid:88)
2
(cid:33)S−1(cid:88)
S−2(cid:88)
S−1(cid:88)
2 ,
+
S−1(cid:88)
2 ,
S−1(cid:88)
1
kj
j=0
1
kj
j=0
2
kjk(cid:96)
(cid:96)=j+1
6
(13)
(14)
2tjt(cid:96)
as in Eqs. 7 and 8 of the main text, where the last line in each case recalls Eq. 5 from the main text.
Analytic minimization of timing variance using Lagrange multipliers
To find the minimum variance when the mean is fixed to be t∗, we utilize Lagrange multipliers. Because the variance
is a function of only the first and second moments and is monotonically increasing with the second moment, finding
the minimum of the variance with a fixed mean is equivalent to finding the minimum of the second moment with a
fixed mean. Thus, the set of r(cid:96) values which produces the minimum variance is the set which solves
(cid:0)(cid:10)t2(cid:11) − λ(cid:104)t(cid:105)(cid:1)
0 =
∂
∂r(cid:96)
(15)
(16)
However, Eq. 15 raises an issue. Assume that x∗ = y∗ = 1.
for Lagrange multiplier λ.
In this case, there are only three possible rates
αxy and βxy, namely α00, β00, and β10. There are also only two possible paths: (cid:126)s1 = [{0, 0} ,{0, 1}] and (cid:126)s2 =
[{0, 0} ,{1, 0} ,{1, 1}]. Putting these rates and paths into Eqs. 13 and 14 yields
(cid:104)t(cid:105) =
β00
α00 + β00
=
1
α00 + β00
1
(cid:18)
α00 + β00
α00
β10
1 +
α00
α00 + β00
β10
β10
+
(cid:19)
(cid:18)
1
α00 + β00
+
1
β10
(cid:19)
and
(cid:10)t2(cid:11) =
=
β00
α00 + β00
2
(α00 + β00)2
2
(cid:18)
(α00 + β00)2 +
1 +
α00
β10
+
(cid:32)
(cid:19)
(α00 + β00)2 +
2
α00
α00 + β00
β10
β10
α00 (α00 + β00)
.
β2
10
(cid:33)
2
(α00 + β00) β10
+
2
β2
10
7
(17)
By putting Eqs. 16 and 17 into Eq. 15 and solving the resulting system of equations, one obtains that some rates must
be negative or even undefined depending on the order in which they are solved. Since negative rates are unphysical,
we can enforce positivity by making the substitutions αxy = exp (axy) /t∗ and βxy = exp (bxy) /t∗ and finding the
minimum variance in (axy, bxy) space rather than (αxy, βxy) space. This procedure can be done without ever leaving
(αxy, βxy) space by noting that ∂/∂a = (∂α/∂a) ∂/∂α = α (∂/∂α) and similarly that ∂/∂b = β (∂/∂β). This allows
Eq. 15 to be rewritten as
(cid:0)(cid:10)t2(cid:11) − λ(cid:104)t(cid:105)(cid:1) .
0 = r(cid:96)
∂
∂r(cid:96)
(18)
(19)
Putting Eqs. 16 and 17 into Eq. 18 yields two possible solutions to the resulting equations: [β00, α00, β10] = [1/t∗, 0, β10]
with σ2 = t2∗ for any value of β10 or [β00, α00, β10] = [0, 2/t∗, 2/t∗] with σ2 = t2∗/2. Of important note is the fact that
when α00 = 0 only the (cid:126)s1 path is available, while when β00 = 0 only the (cid:126)s2 path is available. Thus, the variance is
seen to be extremized when only one possible path is available and all rates along that path are equal. Additionally,
the longer path yields a smaller variance.
This can be seen to be a simple case of a larger trend. For any possible values of x∗ and y∗ it is possible to choose
a set of reaction rates such that there is only one possible path through (x, y) space. When this is done, the product
terms in Eqs. 13 and 14 becomes identically 1 since ri = ki must be true along the one possible path. All other paths
will have ri = 0 for some i and will thus not contribute. This allows Eq. 18 to be easily calculated for any r(cid:96) that is
in the single possible path,
(cid:32)S−1(cid:88)
i=0
(cid:32)S−1(cid:88)
(cid:33)
i=0
(cid:33)
(cid:32)S−1(cid:88)
1
r2
i
i=0
+
1
ri
0 = r(cid:96)
∂
∂r(cid:96)
=
λ
r(cid:96)
− 2
r2
(cid:96)
− 2
r(cid:96)
(cid:33)2
− λ
1
ri
(cid:33)
(cid:32)S−1(cid:88)
i=0
1
ri
Eq. 19 is true for all r(cid:96) along the single path if and only if all r(cid:96) along that path have the same value, which, from
the restriction that the mean first passage time must be t∗ and Eq. 13, means r(cid:96) = S/t∗. Putting these values back
into Eq. 14 then allows the variance to be simply calculated to be σ2 = t∗2/S.
Eq. 18 must hold for all off-path reactions as well. This can be seen to be true by noting that for all other paths at
least one ri must be 0 in the product term. If (cid:96) (cid:54)= i this fact is not changed and that path will still have 0 contribution.
If (cid:96) = i then the r(cid:96) in front of the derivative operator will still force that path to have 0 contribution since no ki can be
0. Similarly, if r(cid:96) is not a reaction that occurs at any state along the one possible path then the derivative will cause
it to vanish since the contribution from the one possible path does not depend on rates that exist in other states,
while if r(cid:96) is a 0 rate that exists at a state in the one possible path then the factor of r(cid:96) in front of the derivative will
cause the whole expression to vanish. Thus, choosing a set of reaction rates such that there is a single possible path
and all rates along that path are equal is a solution to Eq. 18 for all r(cid:96). Additionally, since σ2 = t∗2/S, the longer
that path is the smaller the variance will be. We state this result more generally by establishing three rules which
state that the variance in first passage time is minimized when:
1. Variability in the possible path taken is minimized
2. Rate at which the system moves through state space is as constant as possible
3. The path length through state space is maximized
Derivation of the lower bound on timing variance
If all rates are the same, ki = k, then Eqs. 13 and 14 become
(cid:104)t(cid:105) =
P ((cid:126)s)
(cid:104)S(cid:105)
k
S
k
=
and
We then have
σ2
t(cid:104)t(cid:105)2 =
as in Eq. 10 of the main text.
(cid:88)
(cid:10)t2(cid:11) =
(cid:10)t2(cid:11) − (cid:104)t(cid:105)2
{(cid:126)s}
(cid:104)t(cid:105)2
P ((cid:126)s)
k2 +
=
k2
(cid:104)S(cid:105)2
k2 +
{(cid:126)s}
(cid:88)
(cid:18) S
(cid:32)(cid:104)S(cid:105)
(cid:19)
=
S2
k2
(cid:104)S(cid:105)
k2 +
(cid:10)S2(cid:11)
k2 − (cid:104)S(cid:105)2
k2
(cid:10)S2(cid:11)
(cid:33)
k2 .
=
1
(cid:104)S(cid:105) +
8
(20)
(21)
(22)
σ2
S(cid:104)S(cid:105)2 ,
|
1810.05848 | 1 | 1810 | 2018-10-13T12:00:24 | Traction force microscopy with optimized regularization and automated Bayesian parameter selection for comparing cells | [
"physics.bio-ph",
"q-bio.CB",
"q-bio.QM"
] | Adherent cells exert traction forces on to their environment, which allows them to migrate, to maintain tissue integrity, and to form complex multicellular structures. This traction can be measured in a perturbation-free manner with traction force microscopy (TFM). In TFM, traction is usually calculated via the solution of a linear system, which is complicated by undersampled input data, acquisition noise, and large condition numbers for some methods. Therefore, standard TFM algorithms either employ data filtering or regularization. However, these approaches require a manual selection of filter- or regularization parameters and consequently exhibit a substantial degree of subjectiveness. This shortcoming is particularly serious when cells in different conditions are to be compared because optimal noise suppression needs to be adapted for every situation, which invariably results in systematic errors. Here, we systematically test the performance of new methods from computer vision and Bayesian inference for solving the inverse problem in TFM. We compare two classical schemes, L1- and L2-regularization, with three previously untested schemes, namely Elastic Net regularization, Proximal Gradient Lasso, and Proximal Gradient Elastic Net. Overall, we find that Elastic Net regularization, which combines L1 and L2 regularization, outperforms all other methods with regard to accuracy of traction reconstruction. Next, we develop two methods, Bayesian L2 regularization and Advanced Bayesian L2 regularization, for automatic, optimal L2 regularization. Using artificial data and experimental data, we show that these methods enable robust reconstruction of traction without requiring a difficult selection of regularization parameters specifically for each data set. Thus, Bayesian methods can mitigate the considerable uncertainty inherent in comparing cellular traction forces. | physics.bio-ph | physics |
Traction force microscopy with optimized
regularization and automated Bayesian parameter
selection for comparing cells
Yunfei Huang1, Christoph Schell2, Tobias B. Huber4,5,6, Ahmet Nihat S¸ ims¸ek1, Nils
Hersch3, Rudolf Merkel3, Gerhard Gompper1, and Benedikt Sabass1,*
1Theoretical Soft Matter and Biophysics, Institute of Complex Systems-2 and Institute for Advanced Simulation,
Forschungszentrum Juelich, D-52425, Juelich, Germany
2Institut fur Klinische Pathologie, Universitatsklinikum Freiburg, D-79002, Freiburg, Germany
3Biomechanics, Institute of Complex Systems-7, Forschungszentrum Juelich, D-52425, Juelich, Germany
4Department of Medicine IV, Faculty of Medicine, Medical Center -- University of Freiburg, Germany
5BIOSS Center for Biological Signalling Studies, Albert-Ludwigs-University Freiburg, Germany
6III. Department of Medicine, University Medical Center Hamburg-Eppendorf, Hamburg, Germany
*[email protected]
ABSTRACT
Adherent cells exert traction forces on to their environment, which allows them to migrate, to maintain tissue integrity, and to
form complex multicellular structures during developmental morphogenesis. Traction force microscopy (TFM) enables the
measurement of traction forces on an elastic substrate and thereby provides quantitative information on cellular mechanics
in a perturbation-free fashion. In TFM, traction is usually calculated via the solution of a linear system, which is complicated
by undersampled input data, acquisition noise, and large condition numbers for some methods. Therefore, standard TFM
algorithms either employ data filtering or regularization. However, these approaches require a manual selection of filter- or
regularization parameters and consequently exhibit a substantial degree of subjectiveness. This shortcoming is particularly
serious when cells in different conditions are to be compared because optimal noise suppression needs to be adapted for
every situation, which invariably results in systematic errors.
Here, we systematically test the performance of new methods from computer vision and Bayesian inference for solving the
inverse problem in TFM. We compare two classical schemes, L1- and L2-regularization, with three previously untested schemes,
namely Elastic Net regularization, Proximal Gradient Lasso, and Proximal Gradient Elastic Net. Overall, we find that Elastic Net
regularization, which combines L1 and L2 regularization, outperforms all other methods with regard to accuracy of traction
reconstruction. Next, we develop two methods, Bayesian L2 regularization and Advanced Bayesian L2 regularization, for
automatic, optimal L2 regularization. Using artificial data and experimental data, we show that these methods enable robust
reconstruction of traction without requiring a difficult selection of regularization parameters specifically for each data set. Thus,
Bayesian methods can mitigate the considerable uncertainty inherent in comparing cellular tractions in different conditions.
Introduction
Mechanical forces between cells and their embedding matrix are essential for a variety of biological processes, ranging from
migration of cells -- including immune cells and cancer cells -- to tissue maintenance and organ development, see1 -- 7 for only
a few of the many review articles on this topic. Many of the relevant processes occur on a micrometer, or sub-micrometer
lengthscale, for instance in nascent cell adhesion sites, filopodia, and bacterial adhesion. To understand these processes'
mechanics and its biological control, reliable and accurate methods for measurement of cellular forces are required.
Traction force microscopy (TFM) is a versatile and perturbation-free method yielding a spatial image of substrate stress
exerted by cells on relatively soft elastic gel substrates. This method has its origins in pioneering work by Harris et al.8, who
employed flexible silicone substrates to investigate the mechanical forces that cells generate. Today, traction force microscopy
has become a method that is routinely used in laboratories studying cell biology and soft matter physics around the world.9 -- 14
Traction force microscopy includes three distinct procedures, illustrated in Fig. 1(a): (1) Cells are plated on an elastic substrate
containing fiducial markers allowing to quantify gel deformation visually, for instance fluorescent beads or quantum dots.15, 16
The deformations caused by adherent cells are recorded by taking images of the gel before and after removing the cell. (2) A
discrete gel displacement field u is calculated by tracking the markers. The most common techniques for tracking are particle
tracking velocimetry (PTV) and particle image velocimetry (PIV). (3) Finally, the traction force field f is calculated from the
displacement field u by making use of a mechanical model of the elastic substrate. A variety of methods exist for this purpose,
including finite element methods,17 -- 21 boundary element methods15, 22, 23 and methods operating in Fourier space11, 21, 23 -- 25.
Usually, calculation of traction from displacement requires either filtering or regularization approaches to reduce the effect of
noise. The TFM methodology is limited by two common serious issues that introduce systematic errors. First, the resolution of
the measured traction is usually not high enough to resolve processes at micrometer-sized cellular structures. Secondly, the
most commonly used TFM algorithms require the user to choose a filter or a regularization parameter, which introduces a
considerable degree of subjectivity regarding smoothness and magnitude of the resulting traction. In this article, we suggest
methods for improving the state-of-the art with respect to these issues.
In the standard TFM approach it is assumed that the substrate is a homogeneous, isotropic, and linear elastic half-space.
The mechanical model relating a continuous displacement field Ui(x) to the traction force field Fj(x(cid:48)) on a two-dimensional
(x = (x1,x2)) surface of the gel is expressed as the integral equation26
(cid:90)
Ui(x) =
Gi j(x− x(cid:48))Fj(x(cid:48))d2x(cid:48),
(1)
2
∑
j=1
Ω
where Ω denotes the whole surface of the substrate. The integrand contains a Green's function Gi j(x) = (1 + ν)/(πE)[(1−
ν)δi j/r + νxix j/r3], and E and ν represent the Young modulus and Poisson ratio, respectively. We also write r = x and δi j
is the Kronecker delta function. Calculation of the traction Fj requires inversion of Eq. (1). A very popular and practical
approach is to solve Eq. (1) in Fourier space.23, 24, 27 With this approach, the inversion is often directly feasible if noise in
the displacement data has been filtered prior to calculation of the traction. Optimal filtering, however, requires input of a
prior-defined filter function that imposes a smoothness constraint on the calculated traction. Moreover, spatial clustering of
traction into sparse regions is not conserved when switching from real space to Fourier space. To take advantage of the sparsity
of traction patterns for better reconstruction, one can solve Eq. (1) in real space. Here, the integral in Eq. (1) can be converted
into a matrix product by discretizing the traction field Fj and interpolating it as a piecewise linear, continuous function using
pyramidal shape functions h(x).23, 28 We write the two-dimensional, discrete displacement field as a 2m× 1 vector u, where m
is the number of discretization nodes. The discrete traction field f is a 2n× 1 vector, where n is the number of nodes at which
traction is prescribed. Then, Eq. (1) becomes
u = Mf + s,
(2)
where we also explicitly included the linear acquisition noise s that is present the experimental data. The matrix M represents
the coefficients of a discretized integration. It can be calculated with different techniques. The required convolution of the
Green's function with the shape functions h(x) can be done by numerical integration.23 However, the computational effort can
be significantly reduced by using the convolution theorem in Fourier space on regular grids.28 In the supplementary S1, we
show how this method for fast calculation of M can be extended to irregularly spaced measurements with the help of the shift
theorem. Regardless of the way how M is calculated, the condition number of M is typically very large. This means that even
the smallest noise s leads to very large errors if a direct inversion of Eq. (2) is attempted. A further characteristic of TFM is that
the experimental data is always undersampled, which introduces a large degree of uncertainty. The challenge in traction force
microscopy is to turn Eq. (2) into a well-posed problem, while providing a traction field f with correct magnitude at high spatial
resolution. The standard approaches employed in TFM are L1 or L2 regularization that penalize norms of the solution to render
traction calculation a well-defined numerical problem.
However, linear problems involving large, ill-conditioned matrices occur commonly throughout engineering and physics.
Consequently, a variety of solution methods exist that not yet been employed in the context of TFM. For example, the elastic
net regularization. Literally, the elastic net regularization behaves like a stretchable fishing net that retains "all the big fish"
while removing the small background signal.29 An alternative are proximal gradient methods, which usually operate in wavelet
space and employ adaptive or non-adaptive thresholding of high spatial frequencies.30, 31 Proximal gradient methods are for
instance used for reconstructing lost parts of an image,32 -- 34 for analysis of MRI data,35 and for analysis of genomic data.36, 37
All regularization methods have the weakness that they require selection of one or more constant regularization parameters to
allow discrimination of noise and signal.
Bayesian statistics provides one solution to this problem. From the Bayesian point of view, regularization parameters can be
seen of as random variables that are picked for every experimental sample from a prior distribution. Thus, one infers the most
probable value of the regularization parameters in a conceptually similar way as one infers the solution.38 Such methods do not
require the choice of a regularization parameter and are therefore potentially less prone to subjectiveness than the classical
regularization. The conceptual framework of employing hierarchical priors for data regularization has been exploited for a large
number of different applications, for instance in astrophysics,39 -- 41, machine learning42, mechanical structure monitoring43,
face recognition44, and radar imagery45. Therefore, it is to be expected that Bayesian analysis can be of great use for traction
force reconstruction.
2/20
In this work, we systematically compare a range of different approaches for solving Eq. (2). Altogether, we study the
performance of seven methods, illustrated as a schematic diagram in Fig. 1(b). First, we test various regularization methods.
Among the regularization methods, we complement the classical TFM approaches, L1- and L2 regularization, with previously
untested methods from computer vision, namely Elastic Net (EN) regularization, Proximal Gradient Lasso (PGL), and Proximal
Gradient Elastic Net (PGEN). We find that the new EN regularization scheme has a substantially improved accuracy as
compared to previous approaches but requires considerable extra computational cost. Secondly, we seek to establish Bayesian
models that can automatically perform an optimal regularization of the data. Initial tests indicate that different freely available
Bayesian hierarchical models are of little use for TFM, since the large number of hidden variables, even when used with
sparsity priors, does not enforce sufficient data faithfulness. Instead, we find that the simplest-possible Bayesian models with
global priors yields robust results that can be interpreted as optimal L2 regularization. We study two variants of this algorithm:
Bayesian L2 regularization (BL2), where the magnitude of the noise in the displacement data must be measured separately, and
Advanced Bayesian L2 regularization (ABL2), which requires no extra input. We test the Bayesian methods using artificial data
and real experimental data. Our results suggest that BL2 is not only very robust, but also superior to classical L2 regularization
when measurement noise is large. Most importantly, BL2 automatically determines the degree of regularization, which removes
subjectiveness from the result. This advance is particularly relevant for in-detail comparison of cells in different conditions,
where the varying signal-to-noise ratio previously made an unambiguous comparison challenging.
Methods
Regularization
A common heuristic approach to solve Eq. (2) is regularization of the solution through additional constraints. Here, not only
the residual of (u− Mf) is minimized in a least-squares sense, but also the magnitude of the solution is penalized through
its p-norm denoted by (cid:107)f(cid:107)p. The trade-off between minimization of the residual and minimization of the solution norm is
determined by fixed regularization parameters, λ1 and λ2, leading to a minimization problem of the type
(3)
(cid:2)(cid:107)Mf− u(cid:107)2
2 + λ1(cid:107)R1f(cid:107)1 + λ2(cid:107)R2f(cid:107)2
2
(cid:3).
f = argmin
f
The two norms are explicitly written as x(cid:107)1 = ∑k xk and x(cid:107)2
as the unit matrix I. Of the large number of existing regularization strategies, we will focus on the following:
k. R1 and R2 are functions that are to be defined, e.g.,
2 = ∑k x2
• L2 regularization, employing an L2-norm with λ2 > 0 and λ1 = 0 to penalize traction magnitude through R2 = I is
currently the most common technique used for TFM.23, 46, 47 L2 regularization is also known as ridge regression or
Tikhonov regularization48 and this method efficiently produces a continuous and smooth reconstructed traction field.
This approach conveys a high level of robustness for the inversion problem in real space and also in Fourier space. See
supplementary information for our implementation in real space.49
• L1 regularization, also called Lasso,50 is realized through setting λ2 = 0, λ1 > 0, and R1 = I. With L1 regularization,
small values of the reconstructed signal are efficiently set to zero. L1 regularization is therefore frequently used in the the
field of compressive sensing (CS),51 where the underlying assumption is that the signal can be represented in a sparse
form where all but a few components of the signal vanish. Recently, the technique has found use for TFM28, 52 -- 54 and it
is appropriate for traction fields containing few, sparsely located traction hotspots. In this case, it has been found that L1
regularization improves the ability to distinguish different traction hotspots.
• The elastic net regularization combines L1- and L2 regularization and both parameters through λ1 > 0, λ2 > 0, and
R1 = R2 = I in Eq. (3). Note that the actual implementation also involves a rescaling of the variables as suggested in
Ref.29. This regularization scheme is known to have a better accuracy compared to L1 and L2 regularization if the
coefficient matrix has many, correlated entries.29 EN regularization is well established for a wide variety of applications,
most notably the analysis of genetic data,36, 55 -- 57 but has to date not been used for TFM. For our implementation of the
EN we employ the popular convex optimization solver CVX, see supplementary information.58, 59
• Proximal gradient methods are an alternative approach to the optimization problems arising from the non-differentiable
target functions in L1 regularization (PGL) and the EN regularization (PGEN).30, 60 Here, the penalty terms are chosen to
be wavelet-transforms written as R1f = 2∑l(cid:104)f,Ψl(cid:105) and R2f = ∑l(cid:104)f,Ψl(cid:105), where the Ψl constitute an orthonormal Wavelet
basis.61, 62 The optimization problem is solved through iterative soft thresholding, where the regularization parameters
control the threshold below which the wavelet-coefficients are set to zero. Proximal gradient methods are widely applied
for image inpainting, which is the process of reconstructing lost or deteriorated parts of images.32, 34, 63 -- 65 Therefore,
these methods may be useful for TFM where traction images are reconstructed from undersampled displacement data.
Details regarding our implementation of is given in the supplementary information.
3/20
These schemes have in common that they require the choice of one or two regularization parameters. Selecting the optimal
regularization parameters is often a non-trivial problem. For L2- and L1-regularization, one can use the so-called L-curve
criterion49 to find regularization parameters that provide a tradeoff between minimization of residual from the inverse problem
and the regularization penalty.21, 23, 28, 46, 47, 66 Usually, the regularization parameter is assumed to be located at the inflection
point of a curve described by the norm of the residual versus the norm of the solution in double-logarithmic axes. However, the
L-curve criterion is of limited use for real data, since the inflection point does not always exist. Alternatively, multiple inflection
points can appear, and the points are hard to localize precisely on the employed logarithmic scales. Moreover, the L-curve
criterion does not behave consistently in the asymptotic limit of large system sizes or when the data is strongly corrupted
by noise.67, 68 Hence, in practice, regularization parameters are often chosen by visual inspection of the resulting traction
field. This procedure lacks objectivity and significantly biases any conclusions drawn from later analysis of the traction forces.
Note that this problem is not specific to regularization, but the issue of distinguishing between noise and "real" signal appears
generally with any type of method if the data is processed in any way to reduce noise.
Bayesian approaches for traction reconstruction
The discrete inverse problem u = Mf + s is of two-fold statistical nature. First, the traction varies from cell to cell and from
image to image. Thus, f is a sample drawn from a distribution of possible traction values, which we denote by p(fα) with an
undetermined parameter α. The function p(fα) describes any prior knowledge about the distribution of traction. For reasons
that will become clear below, we will assume that the prior distribution for the 2n× 1 vector f is a Gaussian
p(fα) =
exp[−αEf(f)]
Zf
,
(4)
where Zf = (2π/α)n and Ef = fTf/2. The second source of randomness is the acquisition noise s. Typically, s is assumed to be
drawn from a zero-mean Gaussian with unknown variance 1/β .38, 39, 41, 69 In the language of Bayesian statistics, the probability
distribution p(uf,β ) is called the likelihood function and determines the probability to to measure a particular vector u given a
traction vector f. Since the noise is Gaussian, the likelihood function is
p(uf,β ) =
exp[−β Eu(uf)]
Zu
,
(5)
where Eu(uf) = (Mf− u)T(Mf− u)/2 and Zu = (2π/β )m. m is, as above, the number of two-dimensional displacements. The
likelihood function p(uf,β ) describes a situation that is exactly the reverse of the experimental situation, where we are looking
for the probability of having f given measurements u. This situation is described by the posterior distribution p(fu) and can be
related to the likelihood via Bayes' rule
P(fu,α,β ) =
p(uf,β )p(fα)
p(uα,β )
=
exp[−β Eu(uf)]
Zu
p(fα)
p(uα,β )
.
(6)
2 + α(cid:107)f(cid:107)2
2
(cid:2)β(cid:107)Mf− u(cid:107)2
Here, the marginal likelihood p(uα,β ) is the overall probability of finding the displacements u when the traction distributions
are integrated out. Thus, p(uα,β ) is also called evidence for the model with {α,β ,u}.
Assuming that α and β are known constants, one can maximize the posterior probability P(fu,α,β ) with respect to f.
(cid:3), which is exactly the formula employed for L2
The resulting solution then satisfies fMP = argminf
regularization, Eq. (3), if the parameters α and β are related to the L2 regularization parameter as λ = α/β .70 Thus, our choice
of a Gaussian prior is justified if we intend to perform an L2 regularization. Other popular choices for replacing Eq. (5) as prior
are the Laplace distribution p(fθ ) = (θ /2)exp(−θ /2(cid:107)f(cid:107)1),71 and a product of a Gaussian and a Laplace distribution.72 Using
these priors, we would have found the formulas corresponding to L1 regularization and EN regularization, respectively. Thus,
regularization is equivalent to maximizing the posterior probability of a measurement assuming fixed, known parameters of the
prior distributions.
However, α and β can also be treated as variables whose values can be determined by maximizing their probability
p(α,βu) = p(uα,β )p(α,β )/p(u) ∼ p(uα,β ), where we assume a uniform prior p(α,β ) and we can omit the marginal
probability since it plays no role for the optimization. To calculate the evidence p(uα,β ), we need to integrate out f in the
posterior given in Eq. (6). Due to the Gaussian probabilities, this integration can be done conveniently by expanding the
integrand around the most probable value fMP. On defining K(f) ≡ αEf(f) + β Eu(f) and its Hessian, A ≡ ∇∇K(f), we expand
as K(f) ≈ K(fMP) + (f− fMP)T A(f− fMP)/2. Thus, the evidence becomes
(cid:90)
p(uα,β ) =
p(uf,β )p(fα)d2nf =
1
ZuZf
f
(cid:90)
f
exp[−K(f)]d2nf =
(2π)n(detA)−1/2
ZuZf
exp[−K(fMP)].
(7)
4/20
Taking the logarithm yields the final result
log p(uα,β ) = −αEf(fMP)− β Eu(fMP)− 1
2
log(detA) + nlogα + mlogβ − mlog(2π).
(8)
The right hand side of this equation is a typical example for target functions employed in Bayesian analysis, for example in
the context of data fitting.38 For TFM, numerical calculation of log(detA) requires some care. We employ here a Cholesky
decomposition of the positive matrix A = LLT yielding log(detA) = log(det(LLT)) = 2logΠiLii = 2Σi log(Lii).
The logarithmic evidence, Eq. (8), assumes a maximum at those parameters α and β that are most likely associated
f = 2n − αTrA−1 and
with the measurement u. The implicit equations resulting from maximizing Eq. (8) read38 2 αEMP
2 β Eu = 2m− 2n + αTrA−1. Since fMP and A depend on α and β , these equations need to be solved numerically. Once α and
β are determined, the equivalent optimal L2 regularization parameter follows as λ = α/ β .
We employ two approaches for determining the numerical values of α and β . In the first approach, called Bayesian
L2 regularization (BL2), we estimate the inverse noise variance β directly from the data calculating the variance of the
measured displacements in spatial regions that are very far away from any cell. Thus, in BL2 only α is determined through
maximization of Eq. (8). In the second approach, termed advanced Bayesian L2 regularization (ABL2), we solve directly for α
and β , which result in an increased computational cost. For both approaches, it is imperative to standardize the data to adjust
its spread in different dimensions. For a displacement vector u of length 2m, we first subtract the mean u = u− 12m ¯u with
¯u = ∑2m
i=1 Mi j
i=1(Mi j − ¯Mj)2/(2m− 1))1/2. Thus, we can define a problem matrix where each column is normalized by its
and ω j = (∑2m
spread Mi j = (Mi j − ¯Mj)/ω j. The standardized problem therefore reads ui = Mi j f j, which yields fi = fi/ωi.
i=1 ui/(2m). Next, we calculate the mean and standard deviation for all columns of the matrix M as ¯Mj = 1/(2m)∑2m
Generation of artificial test data
To quantitatively compare the performance of different reconstruction methods, we require artificial data with exactly known
traction force and displacements. The process of generating this data is shown in Fig. 2(a,i)-(a,iv) and involves prescribing
traction force magnitude and direction in distributed circular areas, analytical calculation of the resulting displacements,23
sampling displacements at discrete positions and addition of noise, and finally the reverse traction reconstruction. Supplementary
S5 provides further details on the involved analytical calculations. Throughout the article, artificial test data is generated for gel
substrates with a Young modulus of E = 10 kPa and a Poisson ratio of ν = 0.3. The size of the image plane is arbitrary, but
fixed to 25 µm × 25 µ m and involves 9 or 15 circular traction spots. For these fixed geometries we vary the traction magnitude,
density of displacements, and the noise level.
Evaluation metrics for assessing the quality of traction reconstruction
To evaluate the quality of the reconstructed traction, we introduce four different error measures comparing reconstructed traction
and known original traction. For this purpose, traction at every grid node is written as a two-dimensional vector t = {tx,ty}.
Real traction and reconstructed traction are discriminated by superscripts as treal and trecon. The error measures are calculated
by discriminating traction inside and outside of Ni circular traction patches in a test sample.
• The Deviation of Traction Magnitude at Adhesions (DTMA)23 is defined as
mean j
DTMA =
1
Ni ∑
i
(cid:17)
(cid:16)(cid:107)trecon
(cid:16)(cid:107)treal
(cid:17)
j,i (cid:107)2 −(cid:107)treal
j,i (cid:107)2
j,i (cid:107)2
mean j
,
(9)
where Ni is the number of circular traction patches and the index i runs over all patches. The index j runs over all traction
vectors in one patch. A DTMA of 0 represents a perfect average traction recovery and a negative or positive value implies
underestimation or overestimation, respectively.
• The Deviation of Traction Magnitude in the Background (DTMB) is the normalized difference between the reconstructed
and real traction magnitude outside the circular patches
DTMB =
meank
(cid:0)(cid:107)trecon
k
(cid:1)
(cid:17) ,
k (cid:107)2
(cid:107)2 −(cid:107)treal
j,i (cid:107)2
(cid:16)(cid:107)treal
1
Ni ∑i mean j
(10)
where the index k runs over all traction vectors outside the patches. A DTMB with a magnitude much smaller than unity
implies low background noise in the reconstructed traction.
5/20
• The Signal to Noise Ratio (SNR) for TFM
SNR =
1
Ni ∑i mean j((cid:107)trecon
j,i (cid:107)2)
)
stdk(trecon
k
.
(11)
measures the detectability of a real signal within a noisy background.73 As before, the index k runs over all traction
vectors outside the patches while j is the index of each traction vector in the patch i. The value of the SNR runs from 0 to
infinity where a SNR that is much larger than unity indicates a good separation between traction and noise.
• The Deviation of the traction Maximum at Adhesions (DMA) measures how peak-values of the traction over- or
(cid:105)
underestimate the true value. The quantity is defined as
j,i (cid:107)2)
max j((cid:107)trecon
(cid:104)
j,i (cid:107)2)− max j((cid:107)treal
max j((cid:107)treal
j,i (cid:107)2)
DMA =
1
NA ∑
i
,
(12)
where the maxima of traction magnitude are calculated for each traction patch separately through index j. This error
measure is particularly important since traction maxima are easy to extract from real experimental data. A DMA of 0
means that the local traction maxima in the reconstruction and in the original data are equal. Positive or negative values
of the DMA indicate that the maximum of traction is overestimated or underestimated.
Experimental procedures
Primary murine podocytes were isolated and maintained by following previously published protocols.74 In brief, mGFP
positive podocytes were isolated from mTom/mGFP*Nphs2Cre reporter mice and subsequent FACS based purification resulted
in a primary podocyte culture of highest purity.75 Substratum matrices were prepared according to previously established
protocols.70 Primary podocytes were seeded on polyacrylamide substrates with a Young's modulus of 16 kPA. Before, gels
were functionalized via SULFO-SANPAH based crosslinking of fibronectin (UV light applied for 5 minutes). Cells were
cultivated for 12− 16 hours before imaging. Cover slips were transferred into flow chambers and cell removal to image the
stress-free gel was achieved with a cell micromanipulator (Eppendorf).
Heart muscle cells from embryonic rats were freshly prepared as described previously76. Cover slides were coated with
approximately 70 µm thick silicone elastomer layer produced from a commercial two-component formulation (Sylgard 184,
Dow Corning; mixing ratio 50:1 base to crosslinker by weight; cured overnight at 60◦C). These substrates contained fluorescent
beads in their uppermost layer (FluoSpheres Crimson carboxylate-modified beads; Invitrogen) and were coated with fibronectin
before cell seeding. Details on preparation and cell culture are published elsewhere76, calibration of stiffness77 yielded a Young
module of 15 kPa and a Poisson ratio of 0.5. Live cell microscopy on spontaneously beating cardiac myocytes was performed
and positions of fluorescent beads were determined by cross-correlation as described in detail previously.76, 78
Results
Manual selection of optimal regularization parameters is challenging
The optimal regularization parameters λ1/2 in Eq. (3) are usually unknown. Classical methods for their choice are the L-curve
criterion67, 79 or the generalized cross validation (GCV) for L2 regularization.49, 80 However, these two methods hardly ever
produce the same parameter values and results can differ substantially in the presence of noise, see supporting Fig. S1. To
illustrate the strong effect of regularization on traction reconstruction, we focus on artificial test data where the underlying
traction pattern is known. Figure 2(a) illustrates the generation of artificial traction fields consisting of circular patches each
exerting 100 Pa. Fig. 2c) demonstrates how variation of the regularization parameters affects the error of traction reconstruction
with different methods. Note that the errors exhibit minima for intermediate values of the regularization parameters. For
the methods shown in panels i,iv,v of Fig. 2(c) (L2 regularization, PGL, PGEN), minima occur in the positive error of the
background traction DTMB. In contrast, L1 regularization shown in panel ii of Fig. 2(c) exhibits a maximum in the DTMA,
indicating a parameter regime where the traction magnitude is not underestimated.
The occurrence of clear minima in the error measures suggests that the corresponding regularization parameter values
produce a faithful traction reconstruction. Indeed, employing the values corresponding to the error minima yields traction fields
that visually compare well with the original data, see Figs. 2a) and 2d). Note that for L1 regularization, the reconstruction
clearly overestimates the maximum traction locally. As shown in the supplementary Fig. S2(b), the overestimation of the
maximum quantified by the DMA can only be reduced through ∼ 10 fold reduction of λ1, which however leads to strong
background traction and suppression of real traction, see also Fig. S6. While the minima of the error measures in Fig. 2c) allow
to determine a "best" regularization for test data, the resulting regularization parameter values deviate from those suggested by
6/20
the L-curve criterion, see green lines in Fig. 2c). Moreover, the L-curves for these samples are complex and exhibit multiple
turning points, illustrating the difficulty in choosing the right regularization parameter in experiments, see supporting Fig. S5.
The elastic net outperforms other regularization methods for traction reconstruction
To facilitate quantitative comparison of different reconstruction methods, we employ artificial data consisting of 15 circular
traction spots with traction magnitude between 0 Pa and 250 Pa, see Fig. 3(a). Gaussian noise with a standard deviation given
in percent of the maximal absolute value of the of true displacements is added. The spots have a diameter of 2 µm and the
mesh constant for traction reconstruction is 0.5 µm.
Results from different regularization approaches are shown in Fig. 3(a). The figure illustrates that L2 regularization can
yield realistic estimates for the absolute magnitude of traction on the spots but produces a strong traction background, which
may render identification of traction sites difficult. The opposite deficiencies occur for results from L1 regularization. Here, the
background is nicely suppressed, which can allow excellent resolution of very small traction spots. However, the peak tractions
are significantly overestimated, which can not be mitigated by increasing the regularization parameter, see the supplementary
Fig. S6. Note that the quality of L1 regularization can be improved by using an Iterative Reweighted Least Squares algorithm
and the solution from the L2 regularization as an initial guess, see Supplemental Material. The best results are obtained with the
EN regularization which combines the advantages of L1- and L2-regularization. Here, we obtain a clean background combined
with acceptable accuracy in the absolute traction magnitude on the circular patches. The results from the proximal gradient
methods PGL and PGEN qualitatively have a smooth appearance with a level of background traction that is between those
of L2 and L1. Fig. 3(b) quantifies the described differences between the regularization methods through the error in traction
magnitude on the traction spots (DTMA), the error in traction magnitude in the background (DTMB), signal to noise ratio
(SNR), and error in maximum traction on the spots (DMA). The supplementary Fig. S9 contains additional plots of these
quantities. We find that the reconstruction quality of traction and background improves with increasing number of displacement
measurements m. Furthermore, EN regularization outperforms other regularization methods with regard to reconstruction
accuracy of undersampled data (m/n < 1). However, the advantage of EN regularization comes at a significantly increased
computation time and memory requirement as shown in Table 1.
Reconstruction method
Name
Simulation time
Requirement RAM
Regularization
L2
8 s
L1
75 s
EN
0.8 h
PGL
126 s
PGEN
127 s
Bayesian models
ABL2
BL2
0.1 h
0.5 h
350 MB 1.98 GB 3.87 GB 101 MB 107 MB 400 MB 400 MB
Table 1. Overview of calculation time and RAM requirement for each method. The benchmark results were conducted with a
data set consisting of 1000 displacement measurements and a traction field consisting of 2500 entries.
Bayesian variants of the L2 regularization allow parameter-free traction reconstruction
We next consider the performance of the two Bayesian methods, BL2 and ABL2, that allow automatic choice of the optimal L2
regularization parameter, as schematically shown in Fig. 4(a). Both methods select the optimal regularization parameter by
maximizing the logarithmic evidence, Eq. (8). As illustrated in Fig. 4(a), the regularization parameter is here deduced from the
parameters β and α, characterizing the distributions of measurement noise and traction respectively. We first employ the same
test data as used for Fig. 3, containing 5% Gaussian noise in the displacements with β = 400 Pix−2.
With BL2, the log evidence exhibits a clear maximum in a one-dimensional space as seen in Fig. 4(c). Figure 4(d) shows the
reconstructed traction employing the optimal parameter λ2 = 76.75 Pa2/Pix2. Visual comparison of the color-coded traction
magnitude in Figs. 4(d) and 4(b) clearly shows that the reconstructed traction has the correct range.
For ABL2, the evidence is a function of β and α as seen in Fig. 4(e). Numerical localization of the maximum yields
α = 3.06e4 Pa−2 and β = 394 Pix−2, which is very close to the known input value of β = 400 Pix−2. The optimal regularization
parameter in this case is thus λ2 = α/ β = 77.66 Pa2/Pix2, which agrees well with the estimate from BL2 (76.75 Pa2/Pix2).
The resulting traction map is is shown in Fig. 4(f) and is very similar to the traction map resulting from BL2 in Fig. 4(d). Thus,
BL2 and ABL2 yield consistent parameter estimates that produce traction reconstruction of good accuracy. See supplementary
Fig. S9 for a comparison of the Bayesian methods with non-Bayesian approaches.
As with other regularization approaches, quality of reconstruction strongly depends on the present noise. When the
magnitude of the noise is comparable to the magnitude of the displacements caused by the traction (σn/σ¯u ≈ 1), little
information can be recovered. For instance, the small circular traction sites labeled 1 and 2 in Fig. 3(a) are almost impossible
to detect in the presence of 5% noise, but can be reconstructed in the noise-free case, see the supplementary Fig. S7. To
quantitatively assess the fidelity of reconstruction with small traction forces, we employ a constant 5% but scale the tractions
to mean values of (12 Pa, 16 Pa, 60 Pa, and 120 Pa). The resulting relative strength of noise and displacements is quantified
7/20
through the ratio of standard deviations σn/σ¯u, which is plotted against our reconstruction quality measures in Fig. 4(g). For
comparison, results from manual selection of the regularization parameter using the L-curve criterion are also given. The
reconstruction qualities of BL2, ABL2 and L2 are similar when σn/σ¯u (cid:28) 1 (high traction). However, BL2 and ABL2 have an
improved signal to ratio SNR compared with the L-Curve approach when σn/σ¯u approaches unity, see (iii). This is due to the
difficulties with the L-curve criterion at high noise. The logarithmic evidence function exhibits in all cases a clear maximum,
which enables robust and reliable choice of optimal parameters with BL2 and ABL2. In general, the results from BL2 are
however more reliable since the optimization involves here only one parameter. Overall, the tests with artificial data show
that these Bayesian methods containing few additional parameters to be determined from the data can resolve the ambiguity
associated with manual choice of the regularization parameters over a wide range of signal strengths σn/σ¯u < 1.
BL2 and ABL2 are based on the simplest structure of a Bayesian model with only one, global prior distribution. One
may hypothesize that more complex hierarchies of priors yield an improved traction estimate. For instance, it is possible
to prescribe a position-dependent prior for sparse traction patterns through hierarchical Bayesian networks. Such methods
require more advanced techniques for sampling of the probability distributions and optimization, such as variational techniques
or Markov-chain Monte Carlo methods. We have tested three such algorithms that were originally developed for purposes
other than TFM.71, 81 -- 83 Results are shown in the supplementary information. However, the tested algorithms all produce
highly overestimated, localized traction patterns that sensitively depend on noise. Such errors are likely due to the many free
parameters of the models that, in spite of the sparsity constraints, do not favor a faithful data reconstruction. Thus, our tests
suggests that these hierarchical network models are not suited for the inverse problem associated with TFM.
Test of methods with experimental data
To compare the performance of the different methods for real cells, we employ primary mouse podocytes studied with a
standard TFM setup on polyacrylamide gels having a Young module of ∼ 16 kPa. The deformation field resulting from cellular
traction is shown in Fig. 5 (a). Using this displacement data, we find that the variance of noise is ∼ 0.01pix2 = 103.4nm2 in
regions that are far away from the cell. The maximum displacement is 0.52 µm. Figs. 5(b)-(h) show reconstruction results using
all methods discussed above. As for artificial data, we find here that the EN regularization results in a very clear background.
The traction magnitudes and outlines of adhesion sites are similar to those resulting from regularization with the L2 method.
For L1 regularization, traction localizes in sparse regions and has a significantly higher value than for other methods. Proximal
gradient methods produce smooth traction profiles as expected from the use of the soft wavelet thresholding. The magnitude of
traction measured with PGL and PGEN is close to results of EN and L2.
Next, we considered the performance of the Bayesian methods. The logarithmic evidence, Eq. (8), calculated with BL2 and
ABL2 reveals pronounced maxima, allowing to robustly choose the optimal parameters for the experimental data. See also
supplementary Fig. S11. The resulting values for λ2 are 30.4 Pa2/Pix2 and 24.4 Pa2/Pix2 for BL2 and ABL2, respectively, and
thus agree reasonably well with each other. Figures 5(e),(h) show the traction fields calculated with BL2 and ABL2. These
traction fields are visually very similar to the one obtained with standard L2 regularization. However, the L-curve criterion
provides a much more uncertain estimate of a regularization parameter due to the difficulty of localizing it on a logarithmic
scale, see supplementary Fig. S12. Note that the regularization parameters obtained from the L-curve criterion can not be
directly compared to the parameters resulting from the Bayesian methods due to standardization employed for the latter. Overall,
the suggested Bayesian models can eliminate ambiguity in TFM by automatically providing a consistent parameter choice.
Bayesian regularization enables consistent analysis of traction time sequences
TFM is frequently employed to study dynamical aspects of cell mechanics. Examples include cell migration, cell division, or
cytoskeletal reorganization in response to extracellular stimuli. Such processes are usually accompanied with a change in the
traction distribution. As a result, the optimal regularization parameter varies among different images in a time sequence of
microscopy data. Additionally, the regularization parameter can also change if the degree of noise varies over time, which can
be caused for example by stage drift or photo bleaching. In such cases, it is very challenging to perform a consistent, frame-by
frame analysis to determine the degree of regularization with conventional methods. Thus, one fixed parameter is commonly
employed for the whole time sequence and precision, but also accuracy of traction reconstruction is thereby sacrificed.
To test whether our Bayesian methods can be useful in this situation, we employ TFM data with a spontaneously beating
cardiac myocyte, see Fig. 6. Due to the large size of the cell, we focus on a region of interest shown in Fig. 6(a). The analyzed
time sequence corresponds to one cell contraction. Snapshots from frames 1, 4, and 6 are shown for illustration. Figure 6(b)
shows the overall norm of reconstructed traction where λ2 is either chosen according to the L-curve criterion, held at an
intermediate, constant value, or automatically determined in BL2. The overall traction magnitudes are similar in frames 2-6
where traction is high. Differences occur, however, in the low-traction regime, where BL2 systematically yields lower values of
traction. We expect that the results from BL2 are more trustworthy in this regime since the L-curve criterion yields highly
ambiguous values for the regularization parameters, see supplementary Fig. S13. Figure 6(c) shows the overall norm of the gel
displacement and the optimal regularization parameter estimated with BL2. The noise variance is small, ∼ 0.00003pix2, in
8/20
regions that are far away from the cell (pixel size 0.2 µm). As expected, λBL2 is inversely correlated with the displacement
magnitude.
Figs. 6e show snapshots of the resulting traction fields that illustrate again that BL2 produces slightly different results for
low traction, most apparent in Figs. 6(e)i,I and Fig. 6(f),i. Note that the traction field resulting from classical L2 regularization in
Figs. 6(e),i,I shows a noise background outside of the cell that is almost comparable to the real cellular traction. In contrast, BL2
suppresses this background at the price of an apparently reduced spatial resolution as seen in Fig. 6(f),i. However, this provides
an objective distinction between real signal and noise, which is what is to be expected from a faithful data reconstruction.
Discussion
During the last decade, traction force microscopy has become one of the most popular techniques for studying mechanobiology
on the cellular level. The technique has found broad use for studying sub-cellular structures,66, 84 -- 87 single cell dynamics,88 -- 91
and collective cellular behavior,92 -- 97, and there are far too many applications to be reviewed here appropriately.
Many, if not all, TFM methods critically rely on some form of noise reduction. Usually, traction is calculated from substrate
displacement through the solution of a linear problem involving elastic Green's functions. Here, the effects of noise are not
a technical issue relating to the data precision, but connected directly to the structure of the linear problem where even the
slightest numerical noise can be amplified to an extent that the true solution is entirely lost. The most immediate approach to
deal with noise is to filter the displacement field prior to traction reconstruction. Filtering becomes possible if the solution is
calculated in Fourier space because the convolution theorem simplifies the matrix inversion.24 However, filtering the input data
generally reduces the spatial resolution and optimal resolution can only be gained if the filter is adapted for each sample. In
certain cases, moreover, data filtering is not sufficient to guarantee stability of the solution, for example, if the three-dimensional
position of displacements is included.
A popular alternative strategy for enforcing well-behaved solutions is regularization. With regularization, Fourier-space
inversion becomes more robust.21, 23 However, regularization is also used for real-space approaches and has been used in
conjunction with finite element methods or boundary element methods. Solving the linear problem in real space is generally
more demanding, but has the advantage that the spatial sparsity of traction patterns is conserved. For TFM, two regularization
methods have to date been used, namely L2 regularization23, 70, 73 and L1 regularization28, 53, 54. These methods each have one
regularization parameter that is chosen manually based on heuristics, which introduces a considerable degree of subjectivity in
the resulting traction.
In this work, we systematically compare the classical L1- and L2 regularization to three other methods that have, to our
knowledge, not yet been employed for TFM. These three regularization methods are the Elastic Net (EN), Proximal Gradient
Lasso (PGL) and Proximal Gradient Elastic Net (PGEN). Our tests with artificial data clearly demonstrate that EN regularization
outperforms other regularization methods with regard to the reconstruction quality. Here, accurate traction reconstruction is due
to a simultaneous suppression of background noise and penalizing of large traction magnitude. In contrast, the proximal gradient
methods PGL and PGEN are effective at producing a smoothed traction field, due to the local removal of high-frequency spatial
variations. These results obtained with artificial data agree qualitatively with results from tests with experimental data. Here too,
L1 and L2 regularization yield overestimated or underestimated traction on adhesion sites. EN again yields a clear background
without producing excessively sharp traction peaks at adhesions, see Fig. 5. PGL and PGEN yield smooth traction fields and
rounded adhesion site contours. While our work presents a comprehensive overview of regularization variants in TFM, it does
not cover all variants and solution procedures. For example, it has been suggested to use an L1 norm for both the residual and
regularization term,54 and various other iterative regularization procedures can be tested for TFM in the future.
Next, we ask if Bayesian methods can eliminate the necessity of a manual choice of regularization parameters. Here, the
corresponding parameter values are inferred by maximizing their evidence given a fixed class of chosen probability distributions.
Using the simplest approach, our prior assumption on the traction forces is that they are drawn from one global Gaussian
distribution with an unknown variance 1/α. The posterior distribution determining the probability of a particular traction field
given a measured displacement field is then determined by the parameter 1/α and a further parameter 1/β , quantifying the
variance of the measurement noise. For fixed values of α and β , maximization of the posterior distribution corresponds exactly
to L2 regularization with λ2 given by α/β . However, the values of α and β can also be determined through maximizing their
probability conditioned on a given measurement and the chosen probability distributions. Here, this is equivalent to maximizing
the evidence for u, given any two parameter values. We refer to the simultaneous determination of both parameters from
the evidence as advanced Bayesian L2 regularization (ABL2). In an even simpler approach, we estimate the noise strength
directly from the displacement data, leaving only one parameter α to be determined by maximization of the evidence; which
we call Bayesian L2 regularization (BL2). These methods represent an automatic optimization of the L2 regularization. Thus
the resulting traction field has all the qualitative features of L2 regularization, including the suppression of exceedingly high
traction values and a visible background traction. For all our tests, we found that BL2 was a robust method yielding reasonable
estimates for traction and regularization parameters. Due to the difficulty in choosing the correct regularization parameter
9/20
manually, BL2 has substantial advantages over the classical L2 procedure, in particular if the traction is so small that the
resulting displacements are comparable to the noise σn/σ¯u ≈ 1.
We mention that we have also tested more elaborate hierarchical Bayesian network algorithms that were originally designed
for other purposes than for use with TFM. These include a variational approach termed "Bayesian compressive sensing using
Laplace priors" (BCSL),71 and Markov chain Monte Carlo methods, for instance the "Bayesian Lasso".83 In our experience,
however, none of these methods could compete with the much simpler Bayesian L2 regularization when applied to TFM
problems, see Figs. S8-S10,S14 in the supplementary information.
The advantage of employing Bayesian traction reconstruction is most apparent when cells in different conditions are to
be compared. To perform a correct comparison of situations with different traction, different substrate rigidities, etc., it is
technically necessary to adapt the regularization parameter for each case. However, the difficulty of finding the corresponding
parameters usually makes this impossible, which introduces significant quantitative errors. Bayesian methods present a possible
solution to this problem. We have shown here that BL2 produces regularization that varies smoothly and robustly with strong
temporal changes in the traction exerted by a cell. Thus, we expect that this method can be of wide use for quantitative studies
of cell physiology.
Acknowledgements
C.S. acknowledges support by the Berta-Ottenstein Programme, Faculty of Medicine, University of Freiburg, Germany.
Author contributions statement
Y.H. implemented the algorithms and analyzed the data; N.H. and R.M. performed the experiment on heart muscle cells; C.S.
and T.H. performed the experiment on podocytes; G.G. discussed data and results; B.S. designed and supervised the project and
wrote the main text. All authors participated in writing the manuscript.
Additional information
The authors declare no competing interests.
10/20
References
1. Geiger, B., Bershadsky, A., Pankov, R. & Yamada, K. Transmembrane extracellular matrix-cytoskeleton crosstalk. Nat.
Rev. Mol. Cell Biol. 2, 793 -- 805 (2001).
2. Vogel, V. & Sheetz, M. Local force and geometry sensing regulate cell functions. Nat. Rev. Mol. Cell Biol. 7, 265 -- 275
(2006).
3. Hinz, B. Formation and function of the myofibroblast during tissue repair. J. Invest. Dermatol. 127, 526 -- 537 (2007).
4. Butcher, D. T., Alliston, T. & Weaver, V. M. A tense situation: forcing tumour progression. Nat. Rev. Cancer 9, 108 -- 122
(2009).
5. Califano, J. P. & Reinhart-King, C. A. Exogenous and endogenous force regulation of endothelial cell behavior. J. Biomech.
43, 79 -- 86 (2010).
6. Ringer, P. et al. Multiplexing molecular tension sensors reveals piconewton force gradient across talin-1. Nat. methods 14,
1090 (2017).
7. Lecuit, T., Lenne, P.-F. & Munro, E. Force generation, transmission, and integration during cell and tissue morphogenesis.
Annu. Rev. Cell Dev. Biol. 27, 157 -- 184 (2011).
8. Harris, A. K., Wild, P., Stopak, D. et al. Silicone rubber substrata: a new wrinkle in the study of cell locomotion. Sci. 208,
177 -- 179 (1980).
9. Roy, P., Rajfur, Z., Pomorski, P. & Jacobson, K. Microscope-based techniques to study cell adhesion and migration. Nat.
Cell Biol. 4 (2002).
10. Lange, J. R. & Fabry, B. Cell and tissue mechanics in cell migration. Exp. Cell Res. 319, 2418 -- 2423 (2013).
11. Style, R. W. et al. Traction force microscopy in physics and biology. Soft matter 10, 4047 -- 4055 (2014).
12. Schwarz, U. S. & Soiné, J. R. Traction force microscopy on soft elastic substrates: A guide to recent computational
advances. Biochim. Biophys. Acta 1853, 3095 -- 3104 (2015).
13. Polacheck, W. J. & Chen, C. S. Measuring cell-generated forces: a guide to the available tools. Nat. meth. 13, 415 (2016).
14. Roca-Cusachs, P., Conte, V. & Trepat, X. Quantifying forces in cell biology. Nat. Cell Biol. 19, 742 (2017).
15. Dembo, M. & Wang, Y.-L. Stresses at the cell-to-substrate interface during locomotion of fibroblasts. Biophys. J. 76,
2307 -- 2316 (1999).
16. Bergert, M. et al. Confocal reference free traction force microscopy. Nat. Commun. 7, 12814 (2016).
17. Yang, Z., Lin, J.-S., Chen, J. & Wang, J. H. Determining substrate displacement and cell traction fields -- a new approach.
J. Theo. Biol. 242, 607 -- 616 (2006).
18. Hur, S. S., Zhao, Y., Li, Y.-S., Botvinick, E. & Chien, S. Live cells exert 3-dimensional traction forces on their substrata.
Cell. Mol. Bioeng. 2, 425 -- 436 (2009).
19. Tang, X., Tofangchi, A., Anand, S. V. & Saif, T. A. A novel cell traction force microscopy to study multi-cellular system.
PLoS Comput. Biol. 10, e1003631 (2014).
20. Soiné, J. R. et al. Measuring cellular traction forces on non-planar substrates. Interface focus 6, 20160024 (2016).
21. Kulkarni, A. H., Ghosh, P., Seetharaman, A., Kondaiah, P. & Gundiah, N. Traction cytometry: regularization in the fourier
approach and comparisons with finite element method. Soft matter (2018).
22. Dembo, M., Oliver, T., Ishihara, A. & Jacobson, K. Imaging the traction stresses exerted by locomoting cells with the
elastic substratum method. Biophys. J. 70, 2008 -- 2022 (1996).
23. Sabass, B., Gardel, M. L., Waterman, C. M. & Schwarz, U. S. High resolution traction force microscopy based on
experimental and computational advances. Biophys. J. 94, 207 -- 220 (2008).
24. Butler, J. P., Toli´c-Nørrelykke, I. M., Fabry, B. & Fredberg, J. J. Traction fields, moments, and strain energy that cells exert
on their surroundings. Am. J. Physiol., Cell Physiol. 282, C595 -- C605 (2002).
25. Franck, C., Maskarinec, S. A., Tirrell, D. A. & Ravichandran, G. Three-dimensional traction force microscopy: a new tool
for quantifying cell-matrix interactions. PLoS One 6, e17833 (2011).
26. Landau, L. D. & Lifshitz, E. Theory of elasticity, vol. 7. Course Theor. Phys. 3, 109 (1986).
27. Tolic-Nørrelykke, I. M., Butler, J. P., Chen, J. & Wang, N. Spatial and temporal traction response in human airway smooth
muscle cells. Am. J. Physiol.-Cell Physiol. 283, C1254 -- C1266 (2002).
11/20
28. Han, S. J., Oak, Y., Groisman, A. & Danuser, G. Traction microscopy to identify force modulation in subresolution
adhesions. Nat. Methods 12, 653 -- 656 (2015).
29. Zou, H. & Hastie, T. Regularization and variable selection via the elastic net. J. R. Stat. Soc. Ser. B Stat. Methodol. 67,
301 -- 320 (2005).
30. Combettes, P. L. & Pesquet, J.-C. Proximal splitting methods in signal processing. In Fixed-point algorithms for inverse
problems in science and engineering, 185 -- 212 (Springer, 2011).
31. Schmidt, M., Roux, N. L. & Bach, F. R. Convergence rates of inexact proximal-gradient methods for convex optimization.
In Advances in neural information processing systems, 1458 -- 1466 (2011).
32. Mosci, S., Rosasco, L., Santoro, M., Verri, A. & Villa, S. Solving structured sparsity regularization with proximal methods.
In Joint European Conference on Machine Learning and Knowledge Discovery in Databases, 418 -- 433 (Springer, 2010).
33. Peyré, G., Bougleux, S. & Cohen, L. Non-local regularization of inverse problems. Comput. Vision -- ECCV 2008 57 -- 68
(2008).
34. Fadili, J. M. & Peyré, G. Total variation projection with first order schemes. IEEE Trans. Image Process. 20, 657 -- 669
(2011).
35. Michel, V., Gramfort, A., Varoquaux, G., Eger, E. & Thirion, B. Total variation regularization for fmri-based prediction of
behavior. IEEE Trans. Med. Imaging 30, 1328 -- 1340 (2011).
36. Hannum, G. et al. Genome-wide methylation profiles reveal quantitative views of human aging rates. Mol. cell 49, 359 -- 367
(2013).
37. Sokolov, A., Carlin, D. E., Paull, E. O., Baertsch, R. & Stuart, J. M. Pathway-based genomics prediction using generalized
elastic net. PLoS Comput. Biol. 12, e1004790 (2016).
38. MacKay, D. J. Bayesian interpolation. Neural Comput. 4, 415 -- 447 (1992).
39. Suyu, S. H., Marshall, P., Hobson, M. & Blandford, R. A bayesian analysis of regularized source inversions in gravitational
lensing. Mon. Not. R. Astron. Soc. 371, 983 -- 998 (2006).
40. Vegetti, S. & Koopmans, L. V. Bayesian strong gravitational-lens modelling on adaptive grids: objective detection of mass
substructure in galaxies. Mon. Not. R. Astron. Soc. 392, 945 -- 963 (2009).
41. Ghosh, A., Koopmans, L. V., Chapman, E. & Jeli´c, V. A bayesian analysis of redshifted 21-cm h i signal and foregrounds:
simulations for lofar. Mon. Not. R. Astron. Soc. 452, 1587 -- 1600 (2015).
42. Tipping, M. E. Sparse bayesian learning and the relevance vector machine. J. Mach. Learn. Res. 1, 211 -- 244 (2001).
43. Huang, Y., Beck, J. L., Wu, S. & Li, H. Robust bayesian compressive sensing for signals in structural health monitoring.
Comput.-Aided Civ. Infrastruct. Eng. 29, 160 -- 179 (2014).
44. Qiao, L., Chen, S. & Tan, X. Sparsity preserving projections with applications to face recognition. Pattern Recognit. 43,
331 -- 341 (2010).
45. Zhao, L., Wang, L., Bi, G. & Yang, L. An autofocus technique for high-resolution inverse synthetic aperture radar imagery.
IEEE Trans. Geosci. Remote. Sens. 52, 6392 -- 6403 (2014).
46. Schwarz, U. S. et al. Calculation of forces at focal adhesions from elastic substrate data: the effect of localized force and
the need for regularization. Biophys. J. 83, 1380 -- 1394 (2002).
47. Colin-York, H. et al. Super-resolved traction force microscopy (stfm). Nano Lett. 16, 2633 -- 2638 (2016).
48. Tikhonov, A. N., Arsenin, V. I. & John, F. Solutions of ill-posed problems, vol. 14 (Winston Washington, DC, 1977).
49. Hansen, P. C. Regularization tools version 4.0 for matlab 7.3. Numer. algorithms 46, 189 -- 194 (2007).
50. Tibshirani, R. Regression shrinkage and selection via the lasso. J. Royal Stat. Soc. Ser. B (Methodological) 267 -- 288
(1996).
51. Candès, E. J., Romberg, J. & Tao, T. Robust uncertainty principles: Exact signal reconstruction from highly incomplete
frequency information. IEEE Trans. Inf. Theory 52, 489 -- 509 (2006).
52. Brask, J. B., Singla-Buxarrais, G., Uroz, M., Vincent, R. & Trepat, X. Compressed sensing traction force microscopy. Acta
Biomater. 26, 286 -- 294 (2015).
53. Suñé-Auñón, A., Jorge-Peñas, A., Van Oosterwyck, H. & Muñoz-Barrutia, A. L1-regularized reconstruction for traction
force microscopy. In Biomedical Imaging (ISBI), 2016 IEEE 13th International Symposium on, 140 -- 144 (IEEE, 2016).
12/20
54. Suñé-Auñón, A. et al. Full l 1-regularized traction force microscopy over whole cells. BMC Bioinf. 18, 365 (2017).
55. Sunagawa, S. et al. Structure and function of the global ocean microbiome. Sci. 348, 1261359 (2015).
56. Horlbeck, M. A. et al. Compact and highly active next-generation libraries for crispr-mediated gene repression and
activation. Elife 5, e19760 (2016).
57. Reddy, A. et al. Genetic and functional drivers of diffuse large b cell lymphoma. Cell 171, 481 -- 494 (2017).
58. Grant, M. & Boyd, S. CVX: Matlab software for disciplined convex programming, version 2.1. http://cvxr.com/cvx
(2014).
59. Grant, M. & Boyd, S. Graph implementations for nonsmooth convex programs. In Blondel, V., Boyd, S. & Kimura, H. (eds.)
Recent Advances in Learning and Control, Lecture Notes in Control and Information Sciences, 95 -- 110 (Springer-Verlag
Limited, 2008). http://stanford.edu/~boyd/graph_dcp.html.
60. Parikh, N., Boyd, S. et al. Proximal algorithms. Foundations Trends Optim. 1, 127 -- 239 (2014).
61. Donoho, D. L. & Johnstone, J. M. Ideal spatial adaptation by wavelet shrinkage. Biom. 81, 425 -- 455 (1994).
62. Donoho, D. L., Johnstone, I. M., Kerkyacharian, G. & Picard, D. Wavelet shrinkage: asymptopia? J. Royal Stat. Soc. Ser.
B (Methodological) 301 -- 369 (1995).
63. Figueiredo, M. A. & Nowak, R. D. An em algorithm for wavelet-based image restoration. IEEE Transactions on Image
Process. 12, 906 -- 916 (2003).
64. Beck, A. & Teboulle, M. A fast iterative shrinkage-thresholding algorithm with application to wavelet-based image
deblurring. In ICASSP 2009. IEEE International Conference on Acoustics, Speech and Signal Processing., 693 -- 696 (IEEE,
2009).
65. Peyré, G. The numerical tours of signal processing. Comput. Sci. Eng. 13, 94 -- 97 (2011).
66. Legant, W. R. et al. Multidimensional traction force microscopy reveals out-of-plane rotational moments about focal
adhesions. Proc. Natl. Acad. Sci. U.S.A. 110, 881 -- 886 (2013).
67. Hansen, P. C. The L-curve and its use in the numerical treatment of inverse problems (IMM, Department of Mathematical
Modelling, Technical Universityof Denmark, 1999).
68. Hanke, M. Limitations of the l-curve method in ill-posed problems. BIT Numer. Math. 36, 287 -- 301 (1996).
69. Molina, R., Katsaggelos, A. K. & Mateos, J. Bayesian and regularization methods for hyperparameter estimation in image
restoration. IEEE Trans. Image Process. 8, 231 -- 246 (1999).
70. Plotnikov, S. V., Sabass, B., Schwarz, U. S. & Waterman, C. M. High-resolution traction force microscopy. Methods cell
biology 123, 367 (2014).
71. Babacan, S. D., Molina, R. & Katsaggelos, A. K. Bayesian compressive sensing using laplace priors. IEEE Trans. Image
Process. 19, 53 -- 63 (2010).
72. Li, Q., Lin, N. et al. The bayesian elastic net. Bayesian Analysis 5, 151 -- 170 (2010).
73. Holenstein, C. N., Silvan, U. & Snedeker, J. G. High-resolution traction force microscopy on small focal adhesions-
improved accuracy through optimal marker distribution and optical flow tracking. Sci. Rep. 7, 41633 (2017).
74. Schell, C. et al. N-wasp is required for stabilization of podocyte foot processes. J. Am. Soc. Nephrol. 24, 713 -- 721 (2013).
75. Schell, C. et al. The ferm protein epb41l5 regulates actomyosin contractility and focal adhesion formation to maintain the
kidney filtration barrier. Proc. Nat. Acad. Sci. USA 114, E4621 -- E4630 (2017).
76. Hersch, N. et al. The constant beat: cardiomyocytes adapt their forces by equal contraction upon environmental stiffening.
Biol. Open 2, 351 -- U119 (2013).
77. Cesa, C. M. et al. Micropatterned silicone elastomer substrates for high resolution analysis of cellular force patterns. Rev.
Sci. Instr. 78 (2007).
78. Merkel, R., Kirchgebner, N., Cesa, C. M. & Hoffmann, B. Cell force Microscopy on elastic layers of finite thickness.
Biophys. J. 93, 3314 -- 3323 (2007).
79. Winters, D. W., Van Veen, B. D. & Hagness, S. C. A sparsity regularization approach to the electromagnetic inverse
scattering problem. IEEE Trans. Antennas Propag. 58, 145 -- 154 (2010).
80. Golub, G. H., Heath, M. & Wahba, G. Generalized cross-validation as a method for choosing a good ridge parameter.
Technometrics 21, 215 -- 223 (1979).
13/20
81. Hobert, J. P. & Casella, G. The effect of improper priors on gibbs sampling in hierarchical linear mixed models. J. Am.
Stat. Assoc. 91, 1461 -- 1473 (1996).
82. Tipping, M. E., Faul, A. C. et al. Fast marginal likelihood maximisation for sparse bayesian models. In AISTATS (2003).
83. Korobilis, D. Hierarchical shrinkage priors for dynamic regressions with many predictors. Int. J. Forecast. 29, 43 -- 59
(2013).
84. Balaban, N. Q. et al. Force and focal adhesion assembly: a close relationship studied using elastic micropatterned substrates.
Nat. Cell Biol. 3, 466 (2001).
85. Théry, M., Pépin, A., Dressaire, E., Chen, Y. & Bornens, M. Cell distribution of stress fibres in response to the geometry of
the adhesive environment. Cell Motil. Cytosk. 63, 341 -- 355 (2006).
86. Gardel, M. L. et al. Traction stress in focal adhesions correlates biphasically with actin retrograde flow speed. J. Cell. Biol.
183, 999 -- 1005 (2008).
87. Ray, A. et al. Anisotropic forces from spatially constrained focal adhesions mediate contact guidance directed cell
migration. Nat. Commun. 8 (2017).
88. Chaudhuri, O. et al. Substrate stress relaxation regulates cell spreading. Nat. Commun. 6, 6365 (2015).
89. Valon, L., Marín-Llauradó, A., Wyatt, T., Charras, G. & Trepat, X. Optogenetic control of cellular forces and mechan-
otransduction. Nat. Commun. 8, 14396 (2017).
90. Oria, R. et al. Force loading explains spatial sensing of ligands by cells. Nat. 552, 219 (2017).
91. Sabass, B., Koch, M. D., Liu, G., Stone, H. A. & Shaevitz, J. W. Force generation by groups of migrating bacteria. Proc.
Natl. Acad. Sci. U.S.A. 114, 7266 -- 7271 (2017).
92. Trepat, X. et al. Physical forces during collective cell migration. Nat. Phys. 5, 426 (2009).
93. Mertz, A. F. et al. Cadherin-based intercellular adhesions organize epithelial cell -- matrix traction forces. Proc. Natl. Acad.
Sci. U.S.A. 110, 842 -- 847 (2013).
94. Ng, M. R., Besser, A., Brugge, J. S. & Danuser, G. Mapping the dynamics of force transduction at cell -- cell junctions of
epithelial clusters. Elife 3, e03282 (2014).
95. Steinwachs, J. et al. Three-dimensional force microscopy of cells in biopolymer networks. Nat. Methods 13, 171 (2015).
96. Lembong, J., Sabass, B. & Stone, H. Calcium oscillations in wounded fibroblast monolayers are spatially regulated through
substrate mechanics. Phys. Biol. (2017).
97. Saw, T. B. et al. Topological defects in epithelia govern cell death and extrusion. Nat. 544, 212+ (2017).
14/20
Figure 1. Schematic representation of a typical traction force microscopy (TFM) setup and different reconstruction methods
for TFM. (a) Cells are plated on a planar gel substrate containing fiducial markers. Tracking the markers allows to infer the
deformations u in the surrounding of the cell. These deformations are linearly related to the cellular traction forces f. The
problem of calculating traction f from displacement u is associated with inverting an ill-conditioned matrix M. This problem
can be solved with different reconstruction methods. (b) In this work, we test five regularization methods for traction
reconstruction: L2 regularization (L2), L1 regularization (L1), EN regularization (EN), Proximal Gradient Lasso (PGL) and
Proximal Gradient Elastic Net (PGEN). Furthermore, we develop two Bayesian approaches that do not have any free
parameters, namely Bayesian L2 regularization (BL2) and Advanced Bayesian L2 regularization (ABL2).
15/20
ReconstructionTested reconstruction methodsaRegularization methods: parameter choice requiredBayesian models: no free parameter L2L1ENPGLPGENbBL2Measured substrate displacementsReconstructed tractionABL2Gelufu=MfFigure 2. Systematic tests illustrate substantial ambiguity in the choice of regularization parameters. (a) Schematic of the
employed procedure to test the reconstruction methods. (a,i) Artificial traction pattern consisting of circular spots that
uniformly exert a traction of 100 Pa. (a,ii) Analytical calculation of the gel displacements. (a,iii) The displacement field is
sampled at random positions representing measurements of motion of fiducial markers. (a,iv) Reconstruction of the traction. (b)
Central formula summarizing different regularization approaches. (c) Dependence of various error measures on the
regularization parameters. (b,i)-(b,v) Error measures defined in Eqns. (9)-(12) exhibit various extrema and turning points,
making the definition of an optimal parameter challenging. Note that the minima of the errors do not correspond to values of
regularization parameters suggested by the L-curve criterion (Green dotted lines vs. black dotted lines). DTMA: Deviation of
traction magnitude at adhesion, DTMB: deviation of traction magnitude in background. (b,I)-(b,V) Traction fields calculated
with regularization parameters that correspond to the error minima at the black dotted lines. Space bar: 5 µm.
16/20
Reconstructed traction Traction reconstruction using dierent regularization methods Simulation of measurement Analytical displacementArticial tractioniiiiiIL2L1PGLPGEN 0200[Pa]0200[Pa]022[nm]argmin[Mf-u22+1f1+2f22[fError normsf=Reconstruction worseworseworseworseworseworseworseworse[Pix /Pa ]22[Pix /Pa]2[Pix /Pa ]22[Pix /Pa]2[Pix /Pa]2[Pix /Pa ]22[Pix /Pa]2worseworseIIIVVivivb0200[Pa]aL2:1=02>0,L1:1>02=0,EN:1>02>0,PGL:1>02=0,PGEN:1>02>0,Green dotted line: parameter suggestion from L-curve criterionBlack dotted line: regularization parameter estimated from Error norms in (i)-(v) Eect of regularization parameter for dierent regularization methodsciivENIIIiiiiiFigure 3. The elastic net (EN) outperforms other reconstruction methods in the presence of noise and when applied to
undersampled data. (a) Artificial test data with uniform traction spots and 4% noise in the displacements. Traction maps in
(ii-vi) result from different regularization methods: L2-, L1-, EN, PGL and PGEN, respectively. Space bar 5 µm, displacements
are sampled on average every 0.5 µm (b) Comparison of errors resulting from undersampled data. Undersampling is realized
by reducing the number of displacement vectors m. (c) L-curves with chosen regularization parameters (gray boxes) for a data
set containing 2% noise and m/n = 0.4. (d) Comparison of errors for the regularization parameters shown in (c). EN
regularization shows a favorable tradeoff between error and background signal.
17/20
L2L1ENPGLaPGEN0biiiiiiArticial tractionMean traction errorBackground error( )Signal to noise ratio( )12[Pix /Pa ]22ivMax. traction error( )( )ivcdiL2iiiiviii[Pix /Pa ]22[Pix /Pa ]22[Pix /Pa ]22ENvL1iiiiiPGL[Pix /Pa]2[Pix /Pa]2[Pix /Pa]2ivPGEN[Pa]2500iiiiiiivvviFigure 4. Bayesian L2 regularization (BL2) and Advanced Bayesian L2 regularization (ABL2) are robust methods for
automatic, optimal regularization. (a) Schematic diagram of the procedure employed to infer λ2 in BL2 and ABL2. BL2
requires the variance of the displacement measurements 1/β , which can be obtained by analyzing displacement noise far away
from any cell. ABL2 estimates this noise strength directly from the data. (b) Artificial test data. For the shown results, 5%
Gaussian noise is added to the displacements. Space bars: 5 µm. (c) For BL2, the optimal regularization parameter is located at
the maximum of a one-dimensional plot of the evidence Eq. (8). (d) Reconstruction of traction force using BL2. (e) For ABL2,
the optimal regularization parameter is located at the maximum of a two-dimensional plot of the data evidence vs. α and β .
(f) Reconstruction of traction force using ABL2. (g,i)-(g,iv) Comparison of reconstruction accuracy for L2, BL2 and AbL2.
Different levels of traction forces were applied to change signal-to noise ratio where σn is the standard deviation of the noise
and σ¯u is the standard deviation of the noise-free traction field. Note that BL2 outperforms the other methods for high noise
levels.
18/20
Articial test traction0250[Pa]b0250[Pa]2, ()2ABL2BL2cegiiiiivReconstruction (BL2)Reconstruction (ABL2)argmin[Mf-u22+2f22[ff=max. evid.measure(from )u obtain2= BL2:ABL2:maximize evidence obtain2= and[Pa /Pix ]22[Pa /Pix ]42[1/Pix ]2iidfFigure 5. Test of all reconstruction methods using experimental data. (a) Image of an adherent podocyte with substrate
displacements shown as green vectors. (b)-(h) Reconstructed traction forces using L2, L1, EN, BL2, PGL, PGEN and ABL2,
respectively. Reconstruction with L2-type regularization exhibits a comparatively high background noise. L1-regulation shows
very high, localized traction. Based on tests with artificial data, we expect that these peaks overestimate the traction. The EN
method combines the advantages of L1 and L2 regularization, namely a clean background and localized traction of reasonable
magnitude. PGL and PGEN have smooth traction forces at adhesion and background. (g-h) The Bayesian methods BL2 and
ABL2 yield very similar results as the standard L2 regularization without requiring a search for the optimal regularization
parameters. For better visibility, only every fourth traction vector is shown. Space bar: 25 µm.
19/20
L2PGLPGENL1Cell and gel displacementfcabdeENBL2g2ABL201600[Pa]hm(b, c) L1 regularization(d, e) EN regularization(f, g, h) L2 regularizationFigure 6. Bayesian L2 regularization robustly adapts to different traction levels allowing quantitative analysis of time series.
(a) Image of a spontaneously beating heart muscle cell on a micropatterned substrate allowing to measure displacements.
(b) Overall norm of traction magitude in successive image frames. The maximum corresponds to one contraction of the heart
muscle cell. Traction is calculated with BL2 or, for comparison, via L2 regularization where λ2 is either selected manually for
every frame using the L-curve criterion or held constant throughout the image sequence. (c) Optimal regularization parameter
suggested by BL2 and the norm displacement field correlate. (d,i)-(d,iii) Cell images with displacement field at frames 1, 4, and
6. (e,i)-(e,iii) Snapshots of the traction fields resulting from L2 regularization with a manually chosen parameter λL-curve and a
constant parameter λL-const. in an intermediate range. (f,i)-(f,iii) Snapshots of the traction fields resulting from BL2. Note the
different scaling of displacement and tractions for the different frames. Frame 1 (I) illustrates that BL2 yields a smaller traction
magnitude than L2 in the presence of large noise, where the L-curve criterion is hard to employ. As a result, BL2 allows to
differentiate real traction from noise outside of the cell. For better visibility, only every fourth traction vector is shown.
20/20
Frame=1100 nmL2 regularization2000 nm1000 nmDisplacementsd06[Pa]0120[Pa]060[Pa]20mFrame=4Frame=6BL2efIIIIIIiiiiiibciiiiiiiiiiiia20mCellroi for (b)-(f)[Pa /Pix ]22 |
1606.02191 | 2 | 1606 | 2016-11-28T11:00:20 | Run-and-pause dynamics of cytoskeletal motor proteins | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.SC"
] | Cytoskeletal motor proteins are involved in major intracellular transport processes which are vital for maintaining appropriate cellular function. The motor exhibits distinct states of motility: active motion along filaments, and effectively stationary phase in which it detaches from the filaments and performs passive diffusion in the vicinity of the detachment point due to cytoplasmic crowding. The transition rates between motion and pause phases are asymmetric in general, and considerably affected by changes in environmental conditions which influences the efficiency of cargo delivery to specific targets. By considering the motion of molecular motor on a single filament as well as a dynamic filamentous network, we present an analytical model for the dynamics of self-propelled particles which undergo frequent pause phases. The interplay between motor processivity, structural properties of filamentous network, and transition rates between the two states of motility drastically changes the dynamics: multiple transitions between different types of anomalous diffusive dynamics occur and the crossover time to the asymptotic diffusive or ballistic motion varies by several orders of magnitude. We map out the phase diagrams in the space of transition rates, and address the role of initial conditions of motion on the resulting dynamics. | physics.bio-ph | physics |
Run-and-tumble dynamics of cytoskeletal motor proteins
Anne E. Hafner1,∗, Ludger Santen1, Heiko Rieger1, M. Reza Shaebani1,∗,†
1 Department of Theoretical Physics, Saarland University, 66041 Saarbrücken, Germany
∗ These authors contributed equally to this work.
† Corresponding author. [email protected]
Cytoskeletal motor proteins are involved in major intracellular transport processes which are vi-
tal for maintaining appropriate cellular function. The motor exhibits distinct states of motility:
active motion along filaments, and effectively stationary phase in which it detaches from the fila-
ments and performs passive diffusion in the vicinity of the detachment point due to cytoplasmic
crowding. The transition rates between motion and pause phases are asymmetric in general, and
considerably affected by changes in environmental conditions which influences the efficiency of
cargo delivery to specific targets. By considering the motion of molecular motor on a single fila-
ment as well as a dynamic filamentous network, we present an analytical model for the dynam-
ics of self-propelled particles which undergo frequent pause phases. The interplay between motor
processivity, structural properties of filamentous network, and transition rates between the two
states of motility drastically changes the dynamics: multiple transitions between different types of
anomalous diffusive dynamics occur and the crossover time to the asymptotic diffusive or ballis-
tic motion varies by several orders of magnitude. We map out the phase diagrams in the space
of transition rates, and address the role of initial conditions of motion on the resulting dynamics.
The cellular cytoskeleton is a highly dynamic and complex network of cross-linked biopolymers, which carries out essential
functions in cells such as serving as tracks for motor proteins and adjusting the shape and spatial organization of the
cell. The cytoskeleton enables the cell to efficiently adapt to changes in the environment. Microtubules (MTs) and actin
filaments constitute the dynamic tracks for intracellular transport driven by molecular motors1,2. The cytoskeletal filaments
are polar due to their structural asymmetry, with different spatial organizations. While the orientations of actin filaments
are rather randomly distributed, MTs are usually growing outwards from a microtubule organizing center with their plus
ends facing the cell periphery. MTs often span long distances, of the order of the cell diameter, whereas the mesh size of
the actin cortex is of the order of 100 nm3.
Three different families of motor proteins are involved in the active intracellular transport of organelles and other cargoes:
kinesins, dyneins, and myosins. Molecular motors of each family always move on a specific type of track in a particular
direction4; while kinesins and dyneins usually move along MTs towards the plus and minus ends, respectively, different
types of myosins move on actin filaments towards the plus or minus directions. Motors can attach to the filaments, perform
a number of steps, and detach. Such binding/unbinding cycles are frequently repeated during the motion of motors along
the cytoskeleton. The tendency of the motor to continue its motion along the filament, called motor processivity, varies with
the type of filament and motor5,6, and is highly sensitive to the environmental conditions such as the presence of specific
binding domains or proteins7–9. Active transport on filaments makes the long-distance cargo delivery to specific targets in
cells feasible, whereas the detached phases are extremely inefficient since the cytoplasm is a highly crowded environment
which slows down the transfer of materials10,11. The combination of active and passive motions has been however shown
to be beneficial for optimizing the first-passage properties12.
The motion of molecular motors involves a high degree of complexity due to the dynamics of cytoskeletal filaments,
binding/unbinding cycles of motion, motor-motor interactions, variations of environmental conditions, etc. The dynamics
of cargoes is even more complicated, as they can be transported by teams of motors moving in different directions13. The
analytical studies of the dynamics of molecular motors are often made by adopting simplified schemes, e.g. by ignoring
the chemomechanical cycles and other mechanistic details of stepping. Stochastic two-state models of motion, consisting
of altering phases of active and passive dynamics, have been widely employed to describe the motion of cytoskeletal motor
proteins14,15 and swimming bacteria16–19, and locomotive patterns in other biological systems10,11,20,21. The transition
probabilities between the two states are generally asymmetric and influence the efficiency of cargo delivery18. Interestingly,
bacteria are capable of adjusting the balance between their running and tumbling states in response to the changes in
environmental conditions. For example, it has been recently observed that viscoelasticity of the medium suppresses the
tumbling phase and enhances the swimming speed of E. coli22.
Here, we adopt a coarse-grained approach to study the dynamics of molecular motors on cytoskeleton, which enables us
to identify the impact of motor processivity, structural properties of the filamentous network, and switching frequencies
between the two phases of motion on the transport properties of motors. To isolate the influence of switching frequencies,
we first consider the motion along a single filament and present a random walk model with two states of motility: active
motion along the filament and pausing periods. The latter state emerges due to unbinding from the filament, or due to
waiting periods induced by the presence of obstacles or transport of other motors on the same filament. If unbound from
the filament, the motor practically experiences a subdiffusive dynamics in the crowded cytoplasm and spreads extremely
slowly, thus, it cannot reach far distances and its net displacement is negligible compared to the active motion phase23.
The duration of stay in the pause phase however has a crucial influence on the overall transport dynamics. Hence, we
consider the pauses as a waiting phase when the particle effectively remains stationary along the cytoskeleton until it
continues its active motion again. We focus on the single particle dynamics in this study. Investigation of the motion of a
group of motors is more complicated as, for example, persistent motion in the presence of volume-excluded interactions
between the particles leads to jammed regions or even jamming transition with increasing the particle density19,24–26. The
effects of transition probabilities and volume exclusion on the density and traffic of motors along the filament was recently
studied14. We moreover assume that the transitions between the states of motion occur spontaneously and with asymmetric
rates. Assuming constant transition probabilities in our model results in exponential distributions for the residence times
in each state, which is in agreement with the experimental findings for the distribution of active lifetimes of micron-sized
beads moving along cytoskeleton27. We derive exact analytical expressions for the temporal evolution of the moments of
displacement and show that depending on the transition probabilities between the two states, the motor can experience
crossovers between different anomalous diffusive dynamics on different time scales.
In the second part of this work we study the motion of motor proteins on dynamic cytoskeletal filaments to disentangle the
combined effects of binding/unbinding transition rates, processivity, and structural properties of the underlying network
on transport dynamics. We introduce a coarse-grained perspective to the problem and consider the motion of motors at the
level of traveling between network junctions rather than individual steps along filaments. Since the distribution of direc-
tional change contains rich information about the particle dynamics23,28,29, we characterize the structure of the filamentous
network by probability distributions R(φ) for the angle φ between intersecting filaments, and F (ℓ) for the segment length
ℓ between neighboring intersections. The model consists of two states of motility: active motion and waiting at the junc-
tions. Such waiting periods at junctions have been experimentally observed for transport along cytoskeletal networks6,30,31.
For example, tracking the motion of lysosomes on MT networks has revealed that the particles experience even long pauses
at the nodes of the network before that they can either pass through it or switch to the intersecting filament31. In the mo-
tion state, the motor either moves processively along the previous filament or switches to a new one. The cytoskeleton
is a dynamic network due to the underlying growth and shrinkage of filaments. The dynamics even varies depending
on the cell type and region32. Thus, the transport takes place on a continuously changing structure, which justifies the
relevance of our stochastic approach as the network structure is implicitly given via the probability distributions. Within
the proposed analytical framework, we prove the existence of different regimes of anomalous motion and that the motor
may experience several crossovers between these regimes, as observed in various experiments of motion on cytoskeletal
filaments33–37. It is also shown that the crossover times between different regimes and the asymptotic diffusion coefficient
can vary widely depending on the key parameters of the model: processivity, network structure, and transition rates. We
address the role of initial conditions of motion on the resulting dynamics38, and validate the theoretical predictions by
performing extensive Monte Carlo simulations. The proposed analytical approach and the results are also applicable to
other run-and-tumble motions in biological as well as nonliving systems.
Results
Motion along a single filament
We first consider the motion of a molecular motor on a single filament in a crowded environment. For spontaneously
switching motors, the stochastic motion can be described by two states of motility: (i) ballistic motion along the filament,
and (ii) waiting phase, where the motor remains stationary for a while or unbinds from the filament and performs passive
diffusion until the next binding event happens. When the motor attaches again to the filament, it continues its unidirec-
tional motion. The transition probabilities between the two states of motion are not necessarily symmetric in general, thus,
we consider asymmetric constant transition probabilities κm and κw for the transition from waiting to motion state and
vice versa, respectively [see Fig. 1(A)]. At each time step, the particle either waits or performs a step of length ℓ taken from
a probability distribution F (ℓ). We introduce the probability densities P M
n (x) to find the walker at position x
at time step n in the motion and waiting states, respectively. The temporal evolution of the process can be described by
the following set of coupled master equations
n (x) and P W
P M
n+1(x) =Zdℓ F (ℓ)κm P W
n (x−ℓ)+(1 − κw )P M
n (x−ℓ),
P W
n+1(x) = κw P M
n (x)+(1−κm)P W
n (x).
(1)
We develop a Fourier-z-transform approach39–41, which enables us to obtain exact analytical expressions for the arbitrary
moments of displacement. The details of the theoretical approach can be found in the Suppl. Info., where as an example, the
lengthy expression for the mean squared displacement (MSD) is obtained. It can be seen from Eq.(S17) that, in addition to
the transition probabilities, the results also depend on the initial conditions of motion. In derivation of Eq.(S17), denoting
the probability of initially starting in the motion state by P M
0 , the following initial conditions are imposed
P M
n=0(x) = P M
0 δ(x),
n=0(x) = (1 − P M
P W
0 ) δ(x),
(2)
2
(B)
108
w
m
1
0
0.9 0.1
0.9
0.9
0.1
0.01
0.001
0.001
allistic
b
f u s i v e
f
d i
1
10
102
103
104
105
1
0.8
0.6
0.4
0.2
0
4
10
3
10
c
2
10
10
(A)
(C)
2
1.5
1
0.5
0
1
106
4
10
2
10
1
(D)
ballistic
diffusive
w
m
1
0
0.9 0.1
0.9
0.9
0.1
1
1
1
0.01
1
0.001
1
0.001
10
102
103
104
105
0
0.2
0.4
0.6
0.8
1
FIG. 1 Motion on a single filament. (A) A schematic of the transition probabilities between waiting and motion states of a motor
protein on a single cytoskeletal filament, as described by the set of master equations (1). (B) MSD as a function of the step number
n for λ=1, P M
0 =1, and several values of κm and κw. The solid lines are obtained from the analytic expression (S17) and the symbols
represent Monte Carlo simulation results. (C) Temporal evolution of the anomalous exponent α via Eq. (5), for the same parameters as
in panel (B). (D) The crossover time nc to the asymptotic ballistic motion for λ=1 and P M
0 =1 in the (κw, κm) phase space.
n=0(x)+P W
which results in Pn=0(x)=P M
n=0(x)=δ(x). The analytical results for the time evolution of MSD are shown in
Fig. 1(B) for different values of κm and κw parameters. Several crossovers on different time scales can be observed, even
though the asymptotic behavior is ballistic in all cases as expected from the unidirectional motion of the motor along the
filament. In order to determine the crossovers more quantitatively, we fit the time dependence of the MSD to a power-law
〈x 2〉∼t α, with α being the anomalous exponent. For instance, the following expression can be deduced for the initial
anomalous exponent by fitting the first two steps of motion
α⋆ = ln2 −
2(κw −1)
λ
− κw + κm − 1+
1
0 (−1+κm+κw) ln[2],
κm−P M
(3)
with λ=〈ℓ2〉/〈ℓ〉2 being the variance of the step-length distribution. Assuming a constant step length for simplicity (i.e.
λ=1), if the walker remains in the motion phase forever (i.e. κw =0 and κm=1) one gets α⋆=2 as expected for a ballistic
motion. When starting from the motion state, one has P W
n=0(x)=δ(x), and the initial slope follows
n=0(x)=0 and P M
α⋆ = ln4 − 3κw − κm +
κm
1 − κw ln[2].
(4)
By applying the fit to successive time steps n and n+1, the anomalous exponent as a function of time can be obtained as
α(n) = ln 〈r 2〉n+1
〈r 2〉n ln n + 1
n .
(5)
Figure 1(C) shows the time evolution of the exponent for a motor starting from the motion state. Depending on the
choice of κm and κw rates, the motor may undergo transitions between different regimes of anomalous diffusive dynamics,
e.g. from super- to subdiffusion and then to asymptotic ballistic motion. The crossovers happen on different time scales,
controlled by the transition rates. With increasing κw , the chance of switching to the waiting state grows which, on average,
reduces the initial exponent. On the other hand, smaller values of κm postpone the crossover to the long-term ballistic
3
regime. The combination of these effects may cause up to two crossovers for P M
extrema of α(n) (i.e. if dα/dn=0), or when asymptotically approaching the ballistic regime ( dα
dn
0 =1. We identify a crossover from the
−−→
n→∞
0).
In the limit of n→∞ the terms proportional to n2 dominate and the motility becomes purely ballistic. The asymptotic MSD
is given by
〈x 2〉n→∞ ≈
κ2
m(κm+1)2
(κm+κw )4
〈ℓ〉2 n2.
(6)
The prefactor depends on the switching rates. If the walker never switches to the waiting state, one recovers the relation
〈x 2〉=〈ℓ〉2n2 which is the fastest possible propagation. The crossover to the long-term ballistic regime is approached
asymptotically. We use the distance from the exponent of ballistic motion, i.e. δα = α(n)−2, and estimate the crossover
time as the time step nc at which δα drops below a threshold value ε. Here we report the results for ε=10−2, however, we
checked that the choice of ε does not affect our conclusions. As it can be seen from Fig. 1(D), nc varies by several orders
of magnitude in the (κw , κm) plane. It is expected that nc increases with decreasing κm, as the chance of switching to the
motion state is reduced. The increase of nc at large values of both rates κm and κw is due to frequent state oscillations
which postpone the transition to asymptotic ballistic regime to longer times.
Motion on a dynamic filamentous network
The theoretical framework can be generalized to describe active motion of particles on filamentous networks. We adopt a
coarse-grained approach in which the motion of motor proteins is modeled as a persistent random walk on the intersections
of cytoskeletal filaments. The structure of the network is characterized by the probability distributions R(φ) for the angle
φ between intersecting filaments, and F (ℓ) for the segment length ℓ between neighboring intersections. Furthermore,
a parameter p is introduced to take the processivity of molecular motors into account. The particle either waits at each
time step (waiting state) or walks with a step length ℓ (motion state). In the latter state, the motor either continues along
the previous filament with probability p or chooses a new filament with probability 1−p. Similar to motion on a single
filament, the transition probabilities κm and κw between the two states are assumed to be asymmetric and constant. The
parameter κm (κw ) denotes the switching probability from waiting to motion (motion to waiting) state. The probability
density functions P M
n (x, yθ ) denote the probability to find the walker at position (x, y) along the direction
θ at time step n in the motion and waiting states, respectively. Here a 2D system is considered for simplicity. For extension
of the approach to 3D see28. The evolution of the process is described by the following set of coupled master equations
n (x, yθ ) and P W
P M
n+1(x, yθ ) =
pZ dℓ F (ℓ)κm P W
n x ′, y ′(cid:12)(cid:12)θ+(1−κw ) P M
n x ′, y ′(cid:12)(cid:12)θ+(1−p)Z dℓ F (ℓ)Z π
−π
dγ R(φ)κm P W
n x ′, y ′(cid:12)(cid:12)γ+(1−κw) P M
n x ′, y ′(cid:12)(cid:12)γ,
P W
n+1(x, yθ ) = κw P M
n (x, yθ )+(1−κm)P W
n (x, yθ ).
Assuming isotropic initial conditions
P M
0 (x, yθ ) =
P W
0 (x, yθ ) =
1
2π
1
2π
δ(x)δ( y) P M
0 ,
δ(x)δ( y) (1 − P M
0 ),
(7)
(8)
2π
0 (x, yθ ) = 1
0 (x, yθ ) + P W
which leads to P0(x, yθ ) = P M
δ(x)δ( y), one can follow the proposed Fourier-z-transform
formalism and calculate arbitrary moments of displacement. The exact analytical expression for the MSD, which is the
main quantity of interest, is given in Eq.(S45). The interplay between transition probabilities, motor processivity, initial
conditions of motion, and structure of the underlying network lead to a variety of anomalous transport dynamics on
different time scales. The structure of the network is characterized by the relative variance of F (ℓ) (i.e. λ=〈ℓ2〉/〈ℓ〉2)
and the Fourier transform R of R(φ), which quantify the heterogeneity and anisotropy of the environment, respectively.
R ranges from −1 to 1, with R=0 for a completely random structure, and positive (negative) values for an increased
probability for motion in the near forward (backward) directions. The processivity p and anisotropy R parameters always
appear combined as A = p + R − pR in the solution; thus, by varying A one can separately study the effects of p or R on
the transport of motors. We first choose A=0 in Fig. 2, corresponding to a non-processive motion (p=0) on an isotropic
structure such as actin filament networks (R=0), and then study positive values A=0.5 and A=0.99 in Fig. 3. These
latter choices can e.g. correspond to either a non-processive motion (p=0) on aligned filaments (R>0) such as radially
organized MT networks, or an active motion (p>0) on an isotropic actin network (R=0).
As the phase space of the system is entangled, we restrict ourselves to initially starting from the motion state (P M
0 =1)
in this section, and elaborate on the role of initial conditions in the discussion section. A wide range of different types
of anomalous motion can be observed on varying the transition rates κw and κm. The possible crossovers at short and
4
(C)
κm
1
κw
0
0.9 0.1
0.9
0.9
0.1
0.01
0.001
0.001
allistic
b
f u s i v e
f
d i
10
102
103
104
105
2
1.5
1
0.5
0
1
ballistic
diffusive
κm
1
κw
0
0.9 0.1
0.9
0.9
0.1
1
1
1
0.01
1
0.001
1
0.001
10
102
103
104
105
y/(y+x)
1
0.8
0.6
0.4
0.2
0
(F)
100
= 6
= 2
= 1
10
1
0.001
0.01
0.1
1
A
(A)
(D)
1
0.8
0.6
0.4
0.2
0
(B)
108
106
104
102
1
1
1
0.8
0.6
0.4
0.2
0
(E)
104
103
102
10
0
0.2
0.4
0.6
0.8
1
0
0.2
0.4
0.6
0.8
1
FIG. 2 Non-processive motion on random actin networks. (A) Schematic of the path on a cytoskeletal network during two
consecutive steps of motion, as described by the coupled set of master equations (7). (B) MSD as a function of the step number n for
λ=1, P M
0 =1, A=0, and several values of κm and κw. The solid lines correspond to the analytical expression (S45) and the symbols
represent the simulation results. (C) Temporal evolution of the anomalous exponent α via Eq. (5), for the same parameters as in panel
(B). (D,E) Phase diagrams of (D) the crossover time nc to the asymptotic diffusive regime, and (E) the long-term diffusion constant D∞,
scaled by 1
vs A for several
4
values of λ.
0 =1 in the (κw, κm) plane. (F) The scale parameter Γ = A(λ−2)−λ
v〈ℓ〉Γ , via Eq. (12) for A=0, λ=1, and P M
A−1
(A)
(B)
(C)
108
106
104
2
10
1
1
(D)
108
6
10
4
10
2
10
1
1
w
m
1
0
0.9 0.1
0.9
0.9
0.1
0.01
0.001
0.001
allistic
b
f u s i v e
f
d i
10
102
103
104
105
(E)
allistic
b
f u s i v e
f
d i
10
102
10 3
104
105
2
1.5
1
0.5
0
1
2
1.5
1
0.5
0
1
ballistic
diffusive
w
m
1
0
0.9 0.1
0.9
0.9
0.1
1
.1
1
0.01
1
0.001
1
0.001
10
102
103
104
105
(F)
ballistic
diffusive
10
10 2
103
104
105
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
0
0.2
0.4
0.6
0.8
1
0
0.2
0.4
0.6
0.8
1
4
10
3
10
102
10
5
10
4
10
FIG. 3 Motion on microtubule networks / Processive motion.
In the upper (lower) panels, the results for A=0.5 (A=0.99)
are shown, which correspond to non-processive motion on moderately (highly) aligned filaments or, alternatively, active motion with
moderate (high) processivity on isotropic networks. (A,D) MSD in terms of the step number n for λ=1, P M
0 =1, and several values of
κm and κw. The solid lines (symbols) correspond to analytical results via Eq.(S45) (simulation results). (B,E) Time evolution of the
anomalous exponent α for the same parameter values as in panels (A,D). (C,F) Crossover time nc to the asymptotic diffusive dynamics
in the case of λ=1 and P M
0 =1 in the (κw, κm) phase space.
5
intermediate time scales are even more diverse than those observed for the motion on a single filament. A similar approach
as in the previous section is followed to obtain the time evolution of the anomalous exponent and identify the crossovers.
The long term dynamics is diffusive in all cases as the directional memory is short ranged and the walker eventually gets
randomized on the network. However, we note that the asymptotic diffusive motion might be observable only on time and
length scales which are not accessible in experiments. By fitting to the power-law 〈x 2〉∼t α, the initial anomalous exponent
is obtained as
α⋆ = ln2 −
2(p+R−p R)(κw−1)
λ
− κw + κm − 1+
which in the case of κw =0 and κm=1 reduces to
1
0 (−1+κm+κw ) ln[2],
κm−P M
α⋆ = 1 + ln1 +
p+R−p R
λ
ln[2].
(9)
(10)
For example, one gets α⋆=1 for non-processive motion on random actin networks in the absence of waiting phases, which
is expected to be diffusive on all time scales. From Eq. (9) one can also see the impact of the initial conditions of motion
and heterogeneity of the network on the exponent.
In the long-time limit, the terms linear in n dominate, thus, the motility becomes purely diffusive. From Eq.(S45), the MSD
in the limit of n→∞ is given by
from which the asymptotic diffusion constant can be determined as
〈r 2〉n→∞ ≈
κm
A(λ − 2) − λ
κm + κw
A − 1
〈ℓ〉2 n,
D∞ =
1
4
v〈ℓ〉
κm
A(λ − 2) − λ
κm + κw
A − 1
=
1
4
v〈ℓ〉Γ
κm
κm + κw
,
(11)
(12)
A−1
with Γ = A(λ−2)−λ
and v being the average motor velocity. Figure 2(E) shows how D∞ varies in the space of transition
probabilities. While decreasing κw or increasing κm enhances the diffusion coefficient, changing the rates in the opposite
direction leads to a strong localization. Taking the dependence of Γ on the heterogeneity parameter λ into account [see
Fig. 2(F)], it can be seen that D∞ varies by several orders of magnitude by varying the key parameters of the problem.
Interestingly, the long-term diffusion constant does not depend on the initial conditions.
To estimate the transition time nc to asymptotic diffusive regime, we follow the convergence of the anomalous exponent
towards 1, and determine nc as the earliest time step at which the exponent difference drops below a threshold ε, i.e.
α(n) − 1 < ε.
(13)
The results shown in Figs. 2(D), 3(C), and 3(F), reported for ε=10−2, reveal that nc varies over several orders of magnitude.
On an actin filament network with an average mesh size 〈ℓ〉∼100nm and a typical motor velocity v∼1µm/s, nc ranges
from 0.1 seconds to more than 100 seconds, which might be beyond the possible time window of experiments, thus, all
different regimes of motion are not necessarily realized in practice.
Discussion
In the previous sections we showed that the transport dynamics of molecular motors depend on the initial conditions of
motion. This can be more clearly seen in Figs. 4(A,D), where the MSD for the motion along a single filament is plotted
for a given set of κw and κm parameters and for different probabilities P M
0 of initially starting in the motion state (see
also Figs. 4(B,E) for the influence of initial conditions on the anomalous exponent α). Indeed, the probabilities of finding
the walker in each of the two states of motility at time n, i.e. P M
n , are controlled by the transition rates
κw and κm at long times. Thus, the influence of initial conditions gradually weakens as P M
n converge towards
their stationary values. The sequence of motion and waiting states can be considered as a discrete time Markov chain with
transition probabilities κm and κw . Denoting the initial probabilities by P M
0 , it can be verified that the time
evolution of these probabilities follows
n and P W
n and P W
n =1−P M
0 and P W
0 =1−P M
P M
n =
P W
n =
κm
κm + κw
κw
κm + κw
κm + κw
+ 1 − κm − κwn
− 1 − κm − κwn
κm + κw
κw P M
κw P M
0 − κm(1 − P M
0 − κm(1 − P M
0 ) ,
0 ) .
(14)
6
(B)
(A)
108
106
104
102
1
10-2
a lli s ti c
b
M
P0
1
1
.1
0.5
1
0.1
1
0.01
1
0
ballistic
diffusive
(C)
1
0.8
M
Pn
0.6
W
Pn
0.4
0.2
0
P M
PW
M
P0
1
1
.5
0.5
.1
0.1
0.01
1
0
0
10
102
103
104
105
1
10
102
103
104
105
3
2 .5
2
1.5
1
0.5
0
1
1
10
2
10
3
10
4
10
5
10
(D)
108
106
4
10
102
1
10-2
a lli s ti c
b
(E)
3
2.5
ballistic
2
diffusive
1.5
1
0.5
(F)
1
0.8
M
Pn
0.6
W
Pn
0.4
0.2
0
PW
P M
1
10
102
103
104
5
10
0
1
10
102
103
104
105
1
10
102
103
104
105
FIG. 4 Influence of initial conditions. The dynamics on a single filament in the absence of waiting (κw =0, upper panels) and
for weak switching rates to waiting state (κw=0.1, lower panels) are compared. (A,D) MSD as a function of the step number n for
κm=0.001, λ=1, and different probabilities P M
0 of initially starting in the motion state. (B,E) Temporal evolution of the anomalous
exponent α(n) for the same set of parameters as in panels (A,D). (C,F) Evolution of the Markov chain probabilities P M
n (red lines) and
P W
n (blue lines), via Eqs. (14). The stationary probabilities are given by Eqs. (15).
The probabilities eventually converges to the stationary values
P M
∞ =
P W
∞ =
κm
(κm+κw )
κw
(κm+κw )
,
,
(15)
thus, if the process is initially started with these probabilities, the system is already equilibrated. Otherwise, the dynamics
at short times is influenced by the choice of initial conditions of motion until the system relaxes towards stationary state,
as shown in Figs. 4(C,F).
By varying the probability P M
0 of initially starting in the motion state, we find regimes with an anomalous exponent greater
than 2 in Figs. 4(B,E). To understand the origin of this peculiar behavior we note that the presented MSD is indeed an
ensemble averaged quantity. When starting from the initial probability P M
0 which is smaller (larger) than the equilibrated
value P M
∞ , an acceleration (deceleration) due to the injection of more (less) particles from the waiting to the motion
state occurs. This is especially more obvious for the case of motion along a single filament with P M
0 =κw =0. In this case,
10
10
8
10
6
10
4
10
2
10
1
-2
10
-4
10
1
10
2
10
103
4
10
5
10
FIG. 5 MSD for the motion on a single fila-
ment, starting from the initial condition P M
0 =0
and without any switching from the waiting to
motion state, i.e. κw=0. The results are shown
for λ=1 and several values of κm.
7
all particle are in the waiting state initially, and no transition to the waiting state happens during the process, thus, all
particles switch to the motion state gradually according to the transition probability κm. By injection of more particles
into the ballistic motion state, the exponents greater than 2 appear, as shown in Fig. 5 (see also the supplementary figure
S1). The exponent converges to 2 as the system approaches the stationary state. Figure 5 shows that the initial slope of
the ensemble averaged MSD is nearly 3 for this set of the parameter values, which can be explained as follows. Because
of the constant transition probabilities, the residence time in the waiting state is exponentially distributed. The particles
of the ensemble, which are all in the waiting phase initially, switch to motion at exponentially distributed times t0, i.e.
p(t0)=κme−κm t0 . As soon as a particle switches the state, it performs a ballistic motion with velocity v, thus
and we obtain the ensemble averaged MSD by integration over all possible transition times until time t
r(t) =0,
v(cid:12)(cid:12)t − t0(cid:12)(cid:12) ,
t < t0
t ¾ t0
(16)
(17)
〈r 2(t)〉 =Z t
0
dt0 p(t0)v2(cid:12)(cid:12)t − t0(cid:12)(cid:12)
m 1
3
2
=
v2
κ2
m 2 − 2κm t + κ2
m t 2 − 2e−κm t .
By Taylor expansion around t=0, one finds 〈r 2(t)〉 ≃ v2
κ2
is nearly 3. A similar procedure can be applied to obtain the initial exponents for other sets of parameter values, as those
shown in Figs. 4(B,E).
m t 3 + . . ., which verifies that the initial anomalous exponent
κ3
In summary, we developed a general analytical framework to study the dynamics of molecular motors on cytoskeleton,
which enabled us to identify the influence of filament network structure, binding/unbinding rates, and motor processivity
on the transport properties of motors. The flexibility of our coarse-grained formalism allowed us to consider different
filamentous structures, from a single filament to a complex network of biopolymers characterized by its structural het-
erogeneity and anisotropy. We obtained exact analytical expressions for the arbitrary moments of displacement, and
verified that the motors display a wide range of different types of motion due to the interplay between motor processivity,
structural properties of filamentous network, and transition rates between the two states of motility. One observes that
multiple crossovers occur between different types of anomalous transport at short and intermediate timescales, and that
the crossover time to the asymptotic diffusive or ballistic motion varies by several orders of magnitude. We also addressed
how the initial conditions of motion affect the resulting dynamics.
In the analysis presented in this work, the transitions between the two states of motion were assumed to be spontaneous,
thus, the motion of motors was described by a Markovian process. However, switching between the states might be not
necessarily spontaneous, and the process can be non-Markovian in general, e.g. if the residence time in a state affects its
switching probability to the other state. These generalizations can be handled within our proposed analytical framework
enabling to deal with different types of environments. The analytical formalism is also applicable to study similar dynamics
in other biological as well as nonliving systems, such as the run-and-tumble motion of bacteria in biological media. It is
worthwhile to mention that the motility in the passive state may not be negligible in general. The dynamics in such a case
can be described by altering states of motion with different velocities and run times.
Methods
We develop an analytical Fourier-z-transform formalism to describe the persistent motion of motor proteins along cy-
toskeletal filaments with stochastic pausing periods. A random walk in discrete time and continuous space with two states
of motility is introduced. The walker either performs a persistent motion on filaments or waits when faces obstacles or
tumbles in the crowded cytoplasm. By introducing the probability density functions P M
n (x, yθ ) for the
probability to find the walker at position (x, y) along the direction θ at time step n in the motion or waiting states, the
temporal evolution of the process can be described by a set of coupled master equations, where the transitions from motion
to waiting state and vice versa (denoted by κw and κm, respectively) are assumed to be asymmetric and constant. Then,
arbitrary moments of displacement can be calculated by applying Fourier and z transforms on the set of master equations.
Finally, the quantities of interest can be obtained in the real space and time by means of inverse transforms. A detailed
description of the analytical approach can be found in Suppl Info.
n (x, yθ ) and P W
References
1. Chowdhury, D. Stochastic mechano-chemical kinetics of molecular motors:
A multidisciplinary enterprise from a physicist's perspective.
Phys. Rep. 529, 1-197 (2013).
2. Appert-Rolland, C., Ebbinghaus, M. & Santen L.
Intracellular transport driven by cytoskeletal motors: General mechanisms and defects.
Phys. Rep. 593, 1-59 (2015).
3. Morone, N. et al. Three-dimensional reconstruction of the membrane skeleton at
the plasma membrane interface by electron tomography.
J. Cell Biol. 174, 851-862 (2016).
4. Schliwa, M. & Woehlke, G. Molecular motors. Nature 422, 759-765 (2003).
5. Shiroguchi, K. & Kinosita, K. Myosin V walks by lever action and Brownian motion. Science 316, 1208-1212 (2007).
6. Ali, M. Y. et al. Myosin Va maneuvers through actin intersections and diffuses along microtubules. Proc. Natl. Acad. Sci. USA 104, 4332-4336 (2007).
8
7. Okada, Y., Higuchi, H. & Hirokawa, N. Processivity of the single-headed kinesin KIF1A through biased binding to tubulin. Nature 424, 574-577 (2003).
8. Culver-Hanlon, T. L., Lex, S. A., Stephens, A. D., Quintyne, N. J. & King, S. J. A microtubule-binding domain in dynactin increases dynein processivity
by skating along microtubules. Nat. Cell Biol. 8, 264-270 (2006).
9. Vershinin, M., Carter, B. C., Razafsky, D. S., King, S. J. & Gross, S. P. Multiple-motor based transport and its regulation by Tau.
Proc. Natl. Acad. Sci. USA 104, 87-92 (2007).
10. Höfling, F. & Franosch, T. Anomalous transport in the crowded world of biological cells. Rep. Prog. Phys. 76, 046602 (2013).
11. Bressloff, P. C. & Newby, J. M. Stochastic models of intracellular transport. Rev. Mod. Phys. 85, 135-196 (2013).
12. Benichou, O., Loverdo, C., Moreau, M. & Voituriez, R. Intermittent search strategies. Rev. Mod. Phys. 83, 81-129 (2011).
13. Hancock, W. O. Bidirectional cargo transport: moving beyond tug of war. Nat. Rev. Mol. Cell Biol. 15, 615-628 (2014).
14. Pinkoviezky, I. & Gov, N. S. Transport dynamics of molecular motors that switch between an active and inactive state. Phys. Rev. E 88, 022714 (2013).
15. Thiel, F., Schimansky-Geier, L. & Sokolov, I. M. Anomalous diffusion in run-and-tumble motion. Phys. Rev.E 86, 021117 (2012).
16. Elgeti, J. & Gompper, G. Run-and-tumble dynamics of self-propelled particles in confinement. EPL 109, 58003 (2015).
17. Theves, M., Taktikos,
J., Zaburdaev, V., Stark, H. & Beta, C. A bacterial
swimmer with two alternating speeds of propagation.
Biophys. J. 105, 1915-1924 (2013).
18. Taktikos, J., Stark, H. & Zaburdaev, V. How the motility pattern of bacteria affects their dispersal and chemotaxis. PLoS ONE 8, e81936 (2013).
19. Soto, R. & Golestanian R. Run-and-tumble
exclusion process
environment:
persistent
for
swimmers.
dynamics
in a
crowded
Phys. Rev. E 89, 012706 (2014).
20. Shaebani, M. R., Hafner, A. E. & Santen, L. Geometrical considerations for anomalous transport in complex neuronal dendrites. submitted (2016).
21. Angelani, L. Averaged run-and-tumble walks. EPL 102, 20004 (2013).
22. Patteson, A. E., Gopinath, A., Goulian, M. & Arratia, P. E. Running and tumbling with E. coli in polymeric solutions. Sci. Rep. 5, 15761 (2015).
23. Shaebani, M. R., Sadjadi, Z., Sokolov, I. M., Rieger, H. & Santen L. Anomalous diffusion of self-propelled particles in directed random environments.
Phys. Rev. E 90, 030701(R) (2014).
24. Fily, Y. & Marchetti, M. C. Athermal phase separation of self-propelled particles with no alignment. Phys. Rev. Lett. 108, 235702 (2012).
25. Bialké, J., Löwen, H. & Speck T. Microscopic theory for the phase separation of self-propelled repulsive disks. EPL 103, 30008 (2013).
26. Peruani, F., Deutsch, A. & Bar, M. Nonequilibrium clustering of self-propelled rods. Phys. Rev. E 74, 030904 (2006).
27. Arcizet, D., Meier, B., Sackmann, E., Rädler, J.O. & Heinrich, D. Temporal analysis of active and passive transport
in living cells.
Phys. Rev. Lett. 101, 248103 (2008).
28. Sadjadi, Z., Shaebani, M. R., Rieger, H. & Santen L. Persistent-random-walk approach to anomalous transport of self-propelled particles.
Phys. Rev. E 91, 062715 (2015).
29. Burov, S. et al. Distribution of directional change as a signature of complex dynamics. Proc. Natl. Acad. Sci. USA 110, 19689-19694 (2013).
30. Ross, J. L., Ali, M. Y. & Warshaw, D. M. Cargo transport: molecular motors navigate a complex cytoskeleton. Curr. Opin. Cell Biol. 20, 41-47 (2008).
31. Balint, S., Vilanova, I. V., Alvarez, A. S. & Lakadamyali, M. Correlative live-cell and superresolution microscopy reveals cargo transport dynamics at
microtubule intersections. Proc. Natl. Acad. Sci. USA 110, 3375-3380 (2013).
32. Shaebani, M. R., Pasula, A., Ott, A. & Santen, L. Tracking of plus-ends reveals microtubule functional diversity in different cell types. submitted (2016).
33. Kulic, I. M. et al. The role of microtubule movement in bidirectional organelle transport. Proc. Natl. Acad. Sci. USA 105, 10011-10016 (2008).
34. Caspi, A., Granek, R. & Elbaum, M. Enhanced diffusion in active intracellular transport. Phys. Rev. Lett. 85, 5655-5658 (2000).
35. Caspi, A., Granek, R. & Elbaum, M. Diffusion and directed motion in cellular transport. Phys. Rev. E 66, 011916 (2002).
36. Bruno, L., Levi, V., Brunstein, M. & Desposito, M. A. Transition to superdiffusive behavior in intracellular actin-based transport mediated by molecular
motors. Phys. Rev. E 80, 011912 (2009).
37. Salman, H., Gil,
Y., Granek, R. & Elbaum, M. Microtubules, motor proteins,
and anomalous mean squared displacements.
Chem. Phys. 284, 389-397 (2002).
38. Saxton, M. J. Anomalous diffusion due to binding: a Monte Carlo study. Biophys. J. 70, 1250-1262 (1996).
39. Sadjadi, Z., Miri, M. F., Shaebani, M. R. & Nakhaee, S. Diffusive transport of light in a two-dimensional disordered packing of disks: Analytical approach
to transport mean free path. Phys. Rev. E 78, 031121 (2008).
40. Sadjadi, Z. & Miri, M. F. Diffusive transport of light in two-dimensional granular materials. Phys. Rev. E 84, 051305 (2011).
41. Miri, M. F. & Stark, H.Modelling light transport in dry foams by a coarse-grained persistent random walk. J. Phys. A: Math. Gen. 38, 3743-3749 (2005).
Acknowledgments
This work was funded by the Deutsche Forschungsgemeinschaft (DFG) through Collaborative Research Center SFB 1027
(Projects A3 and A7).
9
|
1011.3470 | 1 | 1011 | 2010-11-15T18:24:03 | Optical tweezer for probing erythrocyte membrane deformability | [
"physics.bio-ph",
"cond-mat.soft"
] | We report that the average rotation speed of optically trapped crenated erythrocytes is direct signature of their membrane deformability. When placed in hypertonic buffer, discocytic erythrocytes are subjected to crenation. The deformation of cells brings in chirality and asymmetry in shape that make them rotate under the scattering force of a linearly polarized optical trap. A change in the deformability of the erythrocytes, due to any internal or environmental factor, affects the rotation speed of the trapped crenated cells. Here we show how the increment in erythrocyte membrane rigidity with adsorption of $Ca^{++}$ ions can be exhibited through this approach. | physics.bio-ph | physics |
APS/123-QED
Optical tweezer for probing erythrocyte membrane deformability
Department of Physics, Indian Institute of Science, Bangalore - 560012, India∗
Manas Khan, Harsh Soni and A.K. Sood1
(Dated: September 20, 2018)
We report that the average rotation speed of optically trapped crenated erythrocytes is direct sig-
nature of their membrane deformability. When placed in hypertonic buffer, discocytic erythrocytes
are subjected to crenation. The deformation of cells brings in chirality and asymmetry in shape that
make them rotate under the scattering force of a linearly polarized optical trap. A change in the
deformability of the erythrocytes, due to any internal or environmental factor, affects the rotation
speed of the trapped crenated cells. Here we show how the increment in erythrocyte membrane
rigidity with adsorption of Ca++ ions can be exhibited through this approach.
PACS numbers:
In mammals, erythrocytes (red blood cells, (RBCs))
play a very critical role in transportation of oxygen and
nutrition via microcirculation. In this process they are
pumped through the capillaries of much smaller cross-
sections. Thus any change in the deformability of the
red blood cells affects the microcirculatory functions and
may cause serious problems. The altered mechanical
properties of the erythrocyte membranes often indicate
the pathogenesis of many diseases [1]. Sickle cell disease
causes the RBCs to be less deformable and more fragile
[2]. Malaria infected red blood cells too show notable
changes in the mechanical properties of the membranes
[1, 3]. Hence, the estimation of the erythrocyte mem-
brane deformability is of paramount importance. Many
single-cell measurement techniques e.g. micropipette as-
piration [4], forced or flow induced stretching of an RBC
with the help of optical tweezers [5 -- 9], microrheology
study using optical magnetic twisting cytometry [10] and
high-frequency electrical deformation tests [11] etc. have
been used for this purpose. Rheological measurements
also have been done on red blood cells suspensions [12 --
15]. Rotations of folded RBCs under circularly polarized
optical trap have been observed too [16]. Here, we re-
port a simpler method to probe the erythrocyte mem-
brane deformability, connecting it to the rotation speed
of crenated RBCs under the scattering force of an optical
trap.
Normal healthy bi-concave shaped (discocytic) ery-
throcytes possess circular symmetry. While in an op-
tical trap a discocytic erythrocyte stands up and aligns
its diameter along the optic axis. Because of its sym-
metry, the trapped RBC does not experience any torque
under the radiation pressure of the trapping laser beam
and therefore it does not show any rotation. However
when the cells are treated with hypertonic buffer, due to
rapid efflux of water they get deformed, losing the inher-
ent symmetry in their shape (crenation) [17]. In most of
the cases the asymmetric deformation brings in a finite
chirality. A non-zero chirality about the optic axis con-
∗ 1 For correspondence : [email protected]
verts the scattering force into a finite torque that makes
the trapped body rotate about the axis [18 -- 20]. Thus
most of the crenated erythrocytes show rotation in an
optical trap [21, 22]. The average speed of rotation de-
pends on the chirality which, in general, increases with
the contortion of the cells.
With the tonicity of the medium, the degree of de-
formation of the RBCs changes and so does the average
rotation speed [22]. On the other hand, if the intrinsic
membrane deformability of the erythrocytes is changed
by any internal or external condition, they would deform
to a different degree albeit being in a similar hypertonic
buffer.
In this letter we report that the average rota-
tion speed of crenated RBCs in an optical trap can be
used as a measure of their membrane deformability. It
is known that the presence of calcium ions (Ca++) en-
hances the lipid ordering in erythrocyte membranes mak-
ing them more rigid [23 -- 25]. We have probed the same
phenomenon through the average rotation speed and ro-
tation probability of optically trapped crenated RBCs in
hypertonic buffer containing different concentrations of
calcium ions.
Fresh blood was obtained from healthy donors.
The erythrocytes were separated by centrifugation and
washed with phosphate buffer saline (PBS). Until the
experiments, the cells were suspended in PBS (isotonic)
and preserved at 4oC. Experiments were done within
36 hours after collecting the fresh blood samples. To
treat the red blood cells with hypertonic buffer of osmo-
larity 1200 mOsm, sodium chloride (N aCl) and calcium
chloride (CaCl2) were mixed in different proportions and
added to the original erythrocyte suspension in PBS (300
mOsm). In this method we obtained 5 different red cell
suspensions in hypertonic buffer mediums of same osmo-
larity 1200 mOsm but with varying calcium ion concen-
trations in them. The relative proportions of N aCl and
CaCl2 in the five different salt mixtures were determined
in such a fashion that in the final suspensions the contri-
bution of the Ca++ ions to the total osmolarity of 1200
mOsm varied from 0% to 25%, where the absolute value
of Ca++ ion concentration changed from 0 to 300 mM .
To rule out the time dependence of the calcium ion treat-
ment on the erythrocytes, each suspension was prepared
and kept for one hour before it was loaded in the sample
cell.
A linearly polarized 1064 nm N d : Y V O4 diode
pumped solid state laser was used to trap the erythro-
cytes. To built the optical trap, the laser beam was
coupled to a modified Axiovert 200 microscope where
a 63X special IR objective focused the beam tightly to
the sample plane. A monochrome digital CCD camera,
with 30f ps capture rate, attached at the binocular port
of the microscope along with an image acquisition card
was utilized to visualize and record the observations. All
the erythrocyte suspensions in different media were ob-
served under the microscope before switching on the trap.
Then the cells in hypertonic buffer solutions having dif-
ferent concentrations of Ca++ ions were trapped at vary-
ing laser powers. To avoid the surface effect the trap was
kept at 25-30 micron above the bottom plate. A total of
75 RBCs from each suspension were trapped at 5 differ-
ent laser powers for better statistics. All the observations
were recorded and stored in computer HDD for off-line
analysis. For each of the suspensions, we counted the
number of cells out of those 75 trapped RBCs that per-
formed rotations under the trap. The rotation speeds
were measured to get their average values as a function
of trapping laser power for all the samples.
(a)
(b)
5 mm
(c)
(d)
FIG. 1: The optical microscope images display erythrocytes
in different conditions. (a) Normal healthy discocytic RBCs
in phosphate buffer saline (300 mOsm). (b) Red blood cells in
hypertonic buffer of osmolarity 1200 mOsm, without calcium
ions. (c) RBCs in hypertonic buffer medium of same osmolar-
ity (1200 mOsm) but containing 150 mM of Ca++ ions. (d)
Erythrocytes in hypertonic buffer (1200 mOsm) consisting of
300 mM of Ca++ ions.
In the isotonic phosphate buffer, before adding the salt
mixtures to make it hypertonic, the cells were biconcave
disk shaped (discocytic) - as expected (Fig. 1(a)). The
erythrocytes in the hypertonic buffer of 1200 mOsm -
without calcium ions, got deformed to a great extent.
The crenated cells lost their disk like symmetric shape
and became fairly contorted. With irregular morphol-
ogy they almost attained the state of acanthocytes (Fig.
1(b)). The other hypertonic media (all were 1200 mOsm)
with calcium ions did not cause that much of distortion
to the symmetric shapes of the erythrocytes. As fluid
came out of the RBCs, they shrank but retained their
2
symmetry to some degree. The smaller and thinner cells
looked more like a disk. The contortions were lesser for
the erythrocytes in media containing higher concentra-
tions of Ca++ ions (Fig. 1(c),(d)).
When trapped, most of the crenated cells started ro-
tating. The RBCs with more deformation rotated faster.
As each of the erythrocyte suspensions contained cells
with a finite distribution in their sizes, ages etc, all the
cells in the sample cell did not behave in the same fash-
ion. Their degree of contortions and the speed of ro-
tations showed a finite spread. Even some of them did
not rotate at all though the others from the same sample
performed rotation at moderate speeds under the trap.
From the hypertonic buffer suspension of RBCs without
any Ca++ ions, 70% of the trapped cells showed rota-
tions. This percentage decreased as the concentration of
the calcium ions in the suspension increased. Not a single
erythrocyte rotated under the trap where the Ca++ ion
concentration was the most, 300 mM . The number frac-
tion of the total trapped crenated cells that performed
rotation and their average rotation speeds are presented
here (Fig. 2) as a measure of the erythrocyte membrane
deformabilities at different calcium ion concentrations.
0 mM Ca++
100 mM Ca++
150 mM Ca++
200 mM Ca++
300 mM Ca++
)
m
p
r
(
d
e
e
p
S
n
o
i
t
a
t
o
R
e
g
a
r
e
v
A
100
80
60
40
20
0
120
160
200
240
280
320
Laser Power ( mW )
FIG. 2: (Color Online) The average rotation speeds of the
trapped crenated erythrocytes have been plotted against the
laser power for five different calcium ion concentrations. For
each concentration, the data points are fitted to a straight
line. The open symbols represent the experimental data
points and the solid lines are the linear fit to them. Typical
rotations of trapped RBCs corresponding to two data points
are shown in the side panel (enhanced online).
For each of the suspensions, the average rotation speed
of the trapped rotors increased linearly with the laser
power, as expected for the asymmetric chiral rotors
[19, 20]. (Two representative rotation observations are
shown in Fig. 2 side panel.) Thus we get five straight
lines for the five hypertonic media containing different
concentrations of calcium ions (Fig. 2). The slopes of
the straight lines decreased monotonically with increas-
ing Ca++ ion concentration, becoming zero for the high-
est calcium ion concentration. The slope (S) that reflects
the intrinsic membrane deformability of the RBCs, has
been plotted against the calcium ion concentrations (Fig.
3(b)). The other parameter that too reflects the defor-
mations of the erythrocytes is the percentage (R) of the
total number of trapped cells which performed rotation.
The percentage, R, decreased with the increasing con-
centration of the calcium ion (Fig. 3(a)), following the
same trend as that of the former.
0
1
0
2
0
3
0
(a)
1/K = 0.1 M
A = 97.20
80
60
40
R
20
(
1/K = 0.1 M
B = 0.63
S
0
5
4
3
2
1
0
0
0
0
0
0
!
0
Concentration of Ca++
0
1
0
2
0
3
( M )
FIG. 3: (Color Online) (a) The bar graph shows R, the per-
centage of the total number of crenated erythrocytes that
rotated under the trap, at varying Ca++ ion concentrations,
C. (b) Slopes (S) of the straight line fits in Fig. 2 are plotted
against C. The dashed curves represent the fit to the data
sets R and S following eqn 1a and 1b respectively.
As the change in deformability of the erythrocytes with
calcium ion concentrations is an effect of the adsorption
3
of the ions to the cell membrane, the variation is expected
to follow the Langmuir isotherm [25]. The isotherm re-
lates the adsorption of molecules or ions (θ, fractional
coverage of the surface) on a surface to the concentra-
tion of the ions (C) as θ = K · C/(1 + K · C), where K is
the Langmuir adsorption constant. In our experiments,
the observables - reflecting the erythrocyte deformability,
drop off with increasing adsorption of the calcium ions.
Therefore, to fit to our experimental data, we use two
complementary equations:
R = Rmax − A · θ (1a) S = Smax − B · θ (1b)
A and B being fitting parameters. Fig. 3 displays the
data sets R and S at varying Ca++ ion concentration
(C) and the corresponding fitting curves as solid lines.
The parameters' values are given in the figure.
To conclude, the fitting of our experimental data shows
that the variation of the average rotation speed of the
trapped crenated RBCs and the probability of their ro-
tation under the trap are the direct signatures of the red
blood cell membrane deformability. The membrane de-
formability of the cells determine the degree of their con-
tortions when subjected to hypertonic media of same os-
molarity, and the contortions become visible through the
rotation speed as well as the probability of rotation under
an optical trap. Albeit being indirect, this method pro-
vides an easier alternative way to investigate the change
of erythrocyte membrane rigidity and stiffness due to any
internal or external abnormalities. Therefore, it could be
used as a generalized diagnostic tool for the diseases that
cause to affect the normal red blood cell membrane de-
formability.
We thank Department of Science and Technology, In-
dia and Council for Scientific and Industrial Research,
India for financial support.
[1] L. H. Miller, D. I. Baruch, K. Marsh and O. K. Doumbo,
[11] H. Engelhardt and E. Sackmann, Biophys. J. 54, 495
Nature 415, 673 (2002).
(1988).
[2] D. E. Discher, N. Mohandas and E. A. Evans, Science
[12] T. Suda, N. Maeda, D. Shimizu, E. Kamitsubo and T.
266, 1032 (1994).
Shiga, Biorheology 19, 555 (1982).
[3] F. K. Glenister, R. L. Copper, A. F. Cowman, N. Mo-
[13] A. Seiyama, Y. Suzuki and N. Maeda, Colloid Polym Sci
handas and B. M. Cooke, Blood 99, 1060 (2002).
271, 63 (1993).
[4] S. Chien, KL. P. Sung, R. Skalak and S. Usami, Biophys.
[14] BD Riquelme, J. Valverde and RJ Rasia, Biorheology 35,
J. 24, 463 (1978).
325 (1998).
[5] K. Svoboda, C. F. Schmidt, D. Branton and S. M. Block,
[15] A. Drochon, European Physical Journal-Applied Physics
Biophys. J. 63, 784 (1992).
22, 155 (2003).
[6] M. Dao, C. T. Lim and S. Suresh, J. Mech. Phys. Solids
51, 2259 (2003).
[7] YP. L. C. Li, KK. Liu and A. C. K. Lai, J. Boimed. Engg.
128, 830 (2006).
[16] J. A. Dharmadhikari, S. Roy, A. K. Dharmadhikari, S.
Sharma and D. Mathur, Appl. Phys. Lett. 85, 6048
(2004).
[17] T. H. Ham, R. F. Dunn, R. W. Sayre and J. R. Murphy,
[8] C. Li and K. K. Liu, J Mater Sci: Mater Med 19, 1529
Blood 32, 847 (1968).
(2008).
[18] P.Galajda and P. Ormos, Appl. Phys. Lett. 78, 249
[9] R. M. Hochmuth, E. A. Evans and D. F. Colvard, Mi-
(2000).
crovasc Res 11, 155 (1976).
[19] M. Khan, A. K. Sood, F. L. Deepak and C. N. R. Rao,
[10] M. PM Marinkovic, K. T. Turner, J. P. Butler, J. J.
Fredberg and S. Suresh, Cell Physiol 293, C597 (2007).
Nanotechnology 17, S287 (2006).
[20] M. Khan, A. K. Sood, F. L. Deepak and C. N. R. Rao,
J. Nanosci. Nanotechnol. 7, 1800 (2007).
[24] E. E. Quist, Biochem. Biophys. Research Comm. 92, 631
[21] M. Khan, S. K. Mohanty and A. K. Sood, Pramana - J.
(1979).
Phys. 65, 777 (2005).
[22] S. K. Mohanty, K. S. Mohanty and P. K. Gupta, Opt.
Express 13, 4745 (2005).
[23] D. R. Anderson, J. L. Davis and K. L. Carraway, J. Biol.
Chem. 252, 6617 (1977).
[25] R. S. Vest, L. J. Gonzales, S. A. Permann, E. Spencer,
L. D. Hansen, A. M. Judd, and J. D. Bell, Biophysical J.
86, 2251 (2004).
4
|
1604.03532 | 1 | 1604 | 2016-04-12T19:19:09 | A Computational Model of YAP/TAZ Mechanosensing | [
"physics.bio-ph",
"q-bio.SC"
] | In cell proliferation, stem cell differentiation, chemoresistance and tissue organization, the ubiquitous role of YAP/TAZ continues to impact our fundamental understanding in numerous physiological and disease systems. YAP/TAZ is an important signaling nexus integrating diverse mechanical and biochemical signals, such as ECM stiffness, adhesion ligand density, or cell-cell contacts, and thus strongly influences cell fate. Recent studies show that YAP/TAZ mechanical sensing is dependent on RhoA-regulated stress fibers. However, current understanding of YAP/TAZ still remains limited due to the unknown interaction between the canonical Hippo pathway and cell tension. To identify the roles of key signaling molecules in mechanical signal sensing and transduction, we present a novel computational model of the YAP/TAZ signaling pathway. This model converts ECM mechanical properties to biochemical signals via adhesion, and integrates intracellular signaling cascades associated with cytoskeleton dynamics. Adhesion molecules, such as FAK, are predicted to rescue YAP/TAZ activity in soft environments via the RhoA pathway. We found that changes of molecule concentrations result in different pattern of YAP/TAZ stiffness response. We also investigate the sensitivity of YAP/TAZ activity to ECM stiffness. In addition, the model shows that the unresolved synergistic effect of YAP/TAZ activity between the mechanosensing and the Hippo pathways can be explained by the interaction of LIMK and LATS. Overall, our model provides a novel platform for studying YAP/TAZ activity in the context of integrating different signaling pathways. This platform can be used to gain new fundamental insights into roles of key molecular and mechanical regulators on development, tissue engineering or tumor progression. | physics.bio-ph | physics | A Computational Model of YAP/TAZ Mechanosensing
Keywords: pathway model, Hippo pathway, synergy, adhesion, cytoskeleton dynamics
Meng Sun1, Fabian Spill2,1,* and Muhammad H. Zaman1,3,*
1. Department of Biomedical Engineering, 44 Cummington Street, Boston MA 02215, USA
2. Department of Mechanical Engineering, Massachusetts Institute of Technology,
77 Massachusetts Avenue, Cambridge, MA 02139, USA
3. Howard Hughes Medical Institute, Boston University, Boston, MA 02215, USA
*Corresponding author
E-mail: [email protected] (MHZ) and [email protected] (FS)
1
Abstract
In cell proliferation, stem cell differentiation, chemoresistance and tissue organization, the ubiquitous role of
YAP/TAZ continues to impact our fundamental understanding in numerous physiological and disease systems.
YAP/TAZ is an important signaling nexus integrating diverse mechanical and biochemical signals, such as
ECM stiffness, adhesion ligand density, or cell-cell contacts, and thus strongly influences cell fate. Recent
studies show that YAP/TAZ mechanical sensing is dependent on RhoA-regulated stress fibers. However, current
understanding of YAP/TAZ still remains limited due to the unknown interaction between the canonical Hippo
pathway and cell tension. Furthermore, the multi-scale relationship connecting adhesion signaling to YAP/TAZ
activity through cytoskeleton dynamics remains poorly understood. To identify the roles of key signaling
molecules in mechanical signal sensing and transduction, we present a novel computational model of the
YAP/TAZ signaling pathway. This model converts ECM mechanical properties to biochemical signals via
adhesion, and integrates intracellular signaling cascades associated with cytoskeleton dynamics. We perform
perturbations of molecular levels and sensitivity analyses to predict how various signaling molecules affect
YAP/TAZ activity. Adhesion molecules, such as FAK, are predicted to rescue YAP/TAZ activity in soft
environments via the RhoA pathway. We also found that changes of molecule concentrations result in different
pattern of YAP/TAZ stiffness response. We also investigate the sensitivity of YAP/TAZ activity to ECM
stiffness, and compare with that of SRF/MAL, which is another important regulator of differentiation. In
addition, the model shows that the unresolved synergistic effect of YAP/TAZ activity between the
mechanosensing and the Hippo pathways can be explained by the interaction of LIMK and LATS. Overall, our
model provides a novel platform for studying YAP/TAZ activity in the context of integrating different signaling
pathways. This platform can be used to gain new fundamental insights into roles of key molecular and
mechanical regulators on development, tissue engineering or tumor progression.
Introduction
One of the fundamental questions of cell biology is how cells organize themselves into complex tissues and
three-dimensional structures. Multiple signals transmitted from neighboring cells and extracellular matrix
(ECM) environment are integrated into intracellular pathways and guide cellular behavior, such as proliferation,
migration and differentiation. Recent years have seen enormous progress in uncovering the roles of the
transcriptional regulator YAP/TAZ, which controls organ size by integrating both physical and biochemical
cues (1–3). First discovered in Drosophila, deregulation of YAP or its paralog TAZ has been found to lead to
massive organ overgrowth (4). YAP/TAZ has been shown to have significant roles in regulating remarkable
biological properties in development, tissue homeostasis and cancer. The activation of the most studied
upstream cascade of YAP/TAZ, the canonical Hippo pathway, is mainly dependent on cell density sensing, and
results in the activation of the core Hippo kinase LATS which regulates YAP/TAZ localization (1). Despite
these ubiquitous roles, the mechanism of YAP/TAZ localization regulation is still poorly understood (2). One of
the models suggests that activated LATS in the Hippo core complex prevents YAP/TAZ from entering the
nucleus via phosphorylation, involved with further cytosolic sequestration through binding with molecules such
as 14-3-3, or degradation (2, 5). YAP/TAZ lacks of a nuclear localization signal (NLS), and the machinery for
their nuclear import/export is unknown (2). Thus it is an important question to investigate which molecules are
involved in this nuclear transport regulation, such as RanGTPases (6). After translocating into the nucleus,
YAP/TAZ acts as a transcriptional co-activator with several transcription factors and thus targets various genes,
including TEAD (2). More recently, attention has been focused on mechanical regulators of YAP/TAZ activity
beyond the core Hippo pathway (7). For instance, increasing stiffness of the external environment (8), or
exerting static stretch (7) has been found to promote YAP/TAZ nuclear translocation and its downstream
transcription activity. Specifically, it was found that these effects are transmitted through RhoA pathways and
resulting stress fiber formation (8). Therefore, YAP/TAZ can be considered as master-integrators of the cellular
2
microenvironment, integrating different cues of mechanical and biochemical nature.
Sensing mechanical properties of ECM alone is a complex process and of fundamental importance in regulating
cell behaviors (9, 10). Stiff or high adhesion environments have been shown to promote both migration and
proliferation (11). For instance, cells have been shown to migrate towards a stiffer ECM environment (durotaxis)
(12) or a surface-bound gradient of ECM adhesion (haptotaxis) (13). Furthermore, manipulation of ECM
stiffness can direct mesenchymal stem cells to differentiate into neuron, myoblasts or osteoblasts in a myosin
II-dependent manner (14). Up-regulating YAP activity on soft substrates is shown to direct mesenchymal stem
cell differentiation towards osteogenesis, which otherwise is observed on stiff substrates (8, 15). Specifically,
the sensing of the mechanical properties of the ECM relies on force-sensitive molecules located at cell-ECM
adhesion complexes, converting the mechanical signals to biochemical signals inside the cell (16, 17). These
adhesion complexes form around integrins, which are membrane-bound heterodimers, and bind to the ECM.
Important downstream targets include FAK (18, 19), which is phosphorylated at tyrosine 397 (20) upon integrin
binding, and Src further enhances FAK phosphorylation sites at other residues to maximize the kinase activity
of FAK (19, 21). Other than adhesion sensing, phosphorylated FAK at tyrosine 397 is also sensitive to the
stiffness of the surrounding matrix in both 2D and 3D environments (11, 22, 23). Importantly, in 3D, FAK
diffuses throughout the cytoplasm but still modulates cell motility by affecting protrusion activity and matrix
deformation (24). 2D substrates have demonstrated that cells increase their assembly of F-actin stress fibers and
focal adhesions via RhoA contractility regulation when substrate rigidity is increased (9, 13, 25). Stiffness
sensing in 3D is also shown to be associated with RhoA dependent cytoskeleton regulation (23, 26).
Mechanical sensing of the microenvironment is a complex and multi-step process with profound implications
for maintaining homeostasis and disease relevance (9, 10, 27, 28). Various quantitative aspects of signal
transmission from the ECM stiffness or adhesion to intracellular biochemical signals remain largely elusive.
While there are several mechanical models (29), and multiple studies of cytoskeleton dynamics, there is a lack
of an integrated model of mechanosensing and YAP/TAZ which enables one to understand the interactions of
various factors influencing YAP/TAZ activity in a quantitative way. Furthermore, such integrated model can be
used to predict the effect of molecular inhibitors or changes in mechanical properties in silico. In this paper, we
aim to bridge this gap through an integrated model that includes mechanical-to-biochemical signal conversion
by adhesion molecules, intracellular signal transmission, cytoskeleton dynamics, and regulation of effectors
relevant to directing transcriptional programs, such as YAP/TAZ (30). We introduce a response function for
YAP/TAZ activity dependent on ECM stiffness, and study how disturbances of molecular activities can alter
this stiffness response function. The model predicts that adhesion regulation has a significant role in regulating
YAP/TAZ activity in mechanical sensing. Changes in FAK shift the stiffness response function horizontally,
such that FAK overexpression rescues YAP/TAZ activity in soft environments, while mDia shifts the stiffness
response function vertically. We also predict that YAP/TAZ is more sensitive than SRF/MAL in response to
changes in ECM properties. Furthermore, our model is able to explain the observed synergistic effect of
mechanical and canonical Hippo regulation on YAP/TAZ activity through incorporating interaction between
LATS and LIMK.
Methods
To understand and predict YAP/TAZ activity, as measured by its amount in the nucleus, we consider a model of
interactions among several key molecules involved in mechanical sensing mechanism (14, 31). Previous
models focused on providing new insights into individual components of mechanical signaling, such as
adhesion (32), Rho GTPase (33), and cytoskeleton dynamics (34–37) separately. However, an integrated
understanding of how adhesion affects cytoskeleton dynamics and stiffness response effector such as YAP/TAZ
is not well characterized. Our model includes key selective molecules from previous models, such as FAK and
3
RhoA, and is extended with the necessary components for YAP/TAZ regulation (8). We model the time
evolution of the concentrations of the molecules per cell by ordinary differential equations. The model
distinguishes between molecules in different cellular compartments, such as YAP/TAZ in the cytoplasm and
nucleus, and F-actin in the cortex and the cytoplasm. It consists of the conversion of mechanical signals to
biochemical signals via adhesion molecules, RhoA GTPase signal transmission, cytoskeleton dynamics, and the
regulation of transcriptional activity (16, 30, 38).
Signaling network
The network of signaling molecules is shown in Figure 1. External ECM properties are felt through adhesion
complexes via integrins. Higher stiffness of ECM correlates with larger protein aggregation at focal adhesions
(39). Thus the clustering of integrins attracts more FAK to activate them (40). For this paper we focus on FAK,
for which there is clear evidence of higher activity in stiffer 3D ECM environments (11, 41). Further
downstream, we include the small GTPase RhoA, which can activate autoinhibited mDia (42, 43) and ROCK
(44) by binding with them. mDia is an actin formin that nucleates actin filaments (45) and accelerates the rate
of actin elongation 5- to 15- fold (46). ROCK enhances myosin activity by phosphorylating myosin light chain
and inhibiting myosin phosphatase (47). It also phosphorylates and activates LIM-kinase, resulting in the
inactivation of the F-actin severing protein cofilin (44). Thus ROCK promotes the assembly of stress fibers and
actomyosin contractility. In turn, increased actin and myosin activity is associated with increased YAP/TAZ
activity (7, 8). In our model these key molecules are sufficient to interpret the mechanism of mechano-sensing
of YAP/TAZ, and the model is formulated in a way that other molecules, such as Rac, Cdc42, profilin, Aip1
(48), Src or capping proteins, can be easily added to the model. The inputs are the ECM mechanical properties
and concentrations of intracellular signaling molecules, while the output is the nuclear translocation of the
relevant transcriptional molecule, YAP/TAZ. The signaling process of another key regulator of differentiation,
SRF/MAL (49), is used in comparison with YAP/TAZ mechanical sensing. The entry of MAL into the nucleus
to interact with SRF transcription factor activity is regulated through monomeric G-actin. G-actin binds to
MAL to prevents it from binding to SRF and thus its downstream transcription (50).
Figure 1 YAP/TAZ mechano-sensing signaling pathway network
This is the scheme of the computational signaling cascade model. The ECM mechanical properties are transmitted to
intracellular signals via adhesion molecules, such as FAK. Adhesion molecules activation induces RhoA binding with GTP.
The RhoA-GTP drives the formation of actomyosin via mDia and ROCK activation. mDia is a formin that polymerizes
F-actin. ROCK can both activate myosin and induce F-actin assembly through LIMK and cofilin. Eventually, the resultant of
F-actin and myosin activity, i.e., stress fibers in 2D systems, leads to the YAP/TAZ nuclear translocation. In comparison
4
with the YAP/TAZ mechanical sensing mechanism, SRF/MAL senses mechanical inputs by being inhibited by G-actin. In
addition, the Hippo core component LATS is sensitive to cell-cell interaction. In our model, the synergistic effect between
the mechanical sensing and the Hippo pathway is predicted due to the interaction of LATS and LIMK.
Conversion from mechanical signals to biochemical signals
ECM stiffness and ligand density are two key ECM properties initially sensed and transmitted to intracellular
signals via adhesion molecules (16). Elucidating the mechanisms of conversion of mechanical signals to
biochemical signals via adhesion molecules is a fundamental question of mechanobiology, and the precise
mechanisms of mechanical sensing in regulating YAP/TAZ activity are poorly understood. Among the large
number of adhesion molecules, we focus on the activity of FAK in our model, since FAK is a key molecule in
adhesion components regulating the downstream signals, such as RhoA GTPase (51). Other components of
adhesion molecules are implicitly incorporated into the model by the usage of a higher-order Hill function,
which represents cooperativity among the groups of adhesion molecules. There is strong evidence showing that
with higher rigidity of ECM, FAK phosphorylation is up-regulated both on 2D substrate and in 3D matrix (11,
22, 23, 41). Over-expression of FAK can rescue cells whose proliferation was inhibited by low adhesion or
reduced cell spreading (52). In 2D, the cells constrained in spreading showed lower FAK phosphorylation than
the spread cells, even when coated with high densities of fibronectin, and had impaired proliferative ability
(52).
Previous modeling approaches have focused on the dependence of the FAK phosphorylation rate on the ECM
ligand density through a Michaelis-Menten function (32, 53). We extend this function by adding the mechanical
properties of ECM, notably its Youngs modulus E
mol . When fitting data regarding the dependence of FAK on
ECM stiffness (11), we found that a second-order Hill function is able to capture the features of the data (Figure
S1 in the Supporting Materials). This is consistent with previous studies showing that a Hill function is
compatible with the high cooperativity of the adhesion molecules such as integrin, talin, etc (51). Hence, the
time dependence of active phosphorylated FAK is given by
dFAK
P(t )
dt
= ksf
( LD(t ) × E
mol(t ))2
C 2 + ( LD(t ) × E
mol(t ))2 ( FAK 0 − FAK
P(t )) − kdf
FAK
P(t ) . (1)
Here, ksf is the maximum activation rate due to the ECM, and the de-phosphorylation rate of FAK is
kdf . C
is the value of LD(t ) × E
mol(t )
when the activation rate of FAK is ksf
/ 2 , that is, half of the maximum
activation rate. We mainly focus on the ECM stiffness in this paper and regard ligand density as constant.
Molecular switch RhoA GTPase
Rho-family GTPases are molecular 'switches' within cells to control the formation and disassembly of actin
cytoskeletal structures, such as stress fibers and filopodia. Adhesion molecules, such as FAK, sense the ECM
and activate Rho-family GTPases. Among them, RhoA is a key GTPase which has been shown to be
up-regulated with increasing rigidity of ECM in 3D environments (11, 41). In detail, phosphorylated FAK
activates RhoA in both a direct (54) and indirect manner, such as via Fyn (55). Hence, we model the
time-dependence of the concentration of RhoA as
dRhoA
GTP
(t )
dt
= k
(γ(FAK
fkρ
P
(t ))2 + 1)(RhoA
0
− RhoA
GTP
(t )) − k
RhoA
dρ
GTP
(t ). (2)
Here, k
fkρ
is the baseline activation rate of RhoA due to the molecules other than phosphorylated FAK, k
dρ
5
is the RhoA-GTP deactivation rate, which is relevant to the hydrolysis rate of RhoA bound GTP to GDP, and
γ is the RhoA activation enhancement/amplification due to active FAK. We have modeled the dependence of
the activation rate of RhoA on phosphorylated FAK as second order, so the steady state
RhoA
GTP
=
k
(γ(FAK
fkρ
P(t ))2 + 1)
k
dρ
+ k
(γ(FAK
fkρ
P(t ))2 + 1)
is a quasi-second-order hill function. This can represent the cooperative process of activating RhoA through
FAK as discussed earlier.
Cytoskeleton regulation
Downstream of RhoA, several cytoskeleton regulators are activated. We focus on ROCK and mDia to model
their time-dependence by
dROCK
A(t )
= k
dt
RhoA
GTP(t )(ROCK
rρ
− ROCK
A(t )) − k
drock
ROCK
A(t ), (3)
0
dmDia
A(t )
= k
dt
RhoA
GTP(t )(mDia
0
mρ
− mDia
A(t )) − k
dmDia
mDia
A(t ) . (4)
Here k
rρ
is the activation rate of ROCK due to RhoA-GTP, k
drock is the deactivation rate of ROCK, k
is
mρ
the activation rate of mDia due to RhoA-GTP, and k
dmDia is the deactivation rate of mDia. Myosin is then
up-regulated by ROCK, and evolves according to
(t )
dMyo
A
dt
= k
mr(εT(ROCK
A(t )) + 1)(Myo
0
− Myo
(t )) − k
Myo
A
dmy
A
(t ) . (5)
Here k
mr is the activation rate of myosin due to pathways other than ROCK, k
is the deactivation rate of
dmy
myosin, and ε is the myosin activation enhancement/amplification due to active ROCK.
Activated ROCK further activates LIM-kinase (LIMK) and LIMK phosphorylates cofilin to prevent its binding
with actin and deactivates its severing and disassembly of F-actin.
dLIMK
A(t )
dt
= k
lr(τT(ROCK
A(t )) + 1)(LIMK
0
− LIMK
A(t )) − k
dl
LIMK
A(t ) , (6)
dCofilin
NP
(t )
dt
= k
turn−over
(Cofilin
0
− Cofilin
NP
(t )) − k
cr(1 − k
LATS
0
ll
)LIMK
A
NP
2(t )Cofilin
(t ). (7)
Here, k
lr is the activation rate of LIMK due to pathways other than ROCK, such as PAK (56), k
dl is the
deactivation rate of LIMK, τ is LIMK activation enhancement/amplification due to active ROCK compared
with the activation rate due to other molecules, k
turn−over is the dephosphorylation rate of cofilin, k
cr is
phosphorylation rate of cofilin due to LIMK, k
ll is the inhibition rate of LIMK-dependent cofilin
phosphorylation due to the total concentration of LATS
0 in the cell, and the function T(ROCK
A(t )) is a
6
smoothing function given by
T(ROCK
A(t )) = smoothing(
0, ROCK
A(t ) ≤ ROCK
B
ROCK
A(t ) − ROCK
B, ROCK
A(t ) > ROCK
B
) .
This T-function approximates the activation starting region of a Michaelis-Menten function. It mimics that
active ROCK concentrations are required to exceed a threshold to activate LIMK. This approximates the flat
region of the deactivating part of the Michaelis-Menten function. Once active ROCK is above the threshold
concentration ROCK
B , the LIMK activation rate is linearly dependent on ROCK
A(t ), approximating the
linear activation part of the Michaelis-Menten function. The smoothing function uses a smooth filter for
connection, so the T-function and its first derivative are continuous. The saturation part of the
Michaelis-Menten function is implicit in the T-function due to the saturation of concentrations of other
molecules, such as FAK and RhoA.
Here we also propose a distinction between cortical F-actin and F-actin mainly in the interior cytoplasm, named
as cytoplasmic F-actin, which is a key component forming stress fibers. The cortical F-actin and cytoplasmic
F-actin have distinct regulators and downstream effectors. In Drosophila, Tsr, which is a cofilin homolog and
accumulates F-actin around the entire cell cortex, does not have an effect in Yki, the YAP homolog, while the
inhibition of capt, another homolog of cofilin, which restricts actin filament accumulation near the apical
surface, promotes Yki activities (57). Cortical actin associating with apical junctions is also latrunculin resistant
(58). Thus, in our model, we distinguish the cortical F-actin and cytoplasmic F-actin. For simplicity, we assume
that the ratio between G-actin and cortical F-actin is constant. The cytoplasmic F-actin is regulated by mDia
and cofilin as follows:
dFcyto(t )
dt
= k
ra(aT(mDia
A(t )) + 1)Gactin(t ) − kdepFcyto(t ) − k fc1cofilinNP(t )Fcyto(t ) . (8)
Here, k
ra is the polymerization rate of cytoplasmic F-actin from G-actin, kdep is the depolymerization rate of
cytoplasmic F-actin, α is the amplification/enhancement of the polymerization rate due to mDia, and k fc1 is
the severing and disassembly rate of cytoplasmic F-actin due to active cofilin. Cofilin is an F-actin assembly
regulator and its effect is dependent on its concentration in vitro: if the ratio of cofilin/actin subunits in a
filament is low (<1/100), this results in persistent filament severing and acceleration of F-actin
depolymerization, while if the ratio is above 1/10, cofilin stabilizes F-actin in a twisted conformation and can
nucleate filaments (56). The severing effect of cofilin itself can be biphasic when interacting with other
cytoskeleton regulators such as Arp2/3 (35). Therefore in our model, cofilin's effect is to reduce the
cytoplasmic F-actin concentration in the range of given endogenous cofilin and actin amounts (36). For
simplicity, we do not explicitly take cortical actin into account and assume it is proportional to G-actin with a
ratio β .
YAP/TAZ dynamics
YAP/TAZ is regulated by both stress fiber/RhoA activity and cell-cell contacts/LATS activity (2). It translocates
to the nucleus to bind with transcription factors such as TEAD and activate them. Phosphorylation by LATS or,
potentially through some currently unknown kinase whose activity is regulated by cytoskeletal tension, leads to
sequestration of YAP/TAZ. We thus model nuclear YAP/TAZ concentrations by
7
dYAP / TAZ N(t )
dt
= (k
CN
+ kCY Fcyton(t )myoA(t ))(YAP / TAZ0
− YAP / TAZ N(t ))
(9)
− kNCYAP / TAZ N(t ) − klyLATSP(t )YAP / TAZ N(t ).
Here, k
CN is the rate of YAP/TAZ trans-locating from the cytoplasm to the nucleus with no active cytoplasmic
F-actin and myosin, k
NC is the rate of YAP/TAZ translocating from nucleus to cytoplasm with no active LATS,
and k
CY is the YAP/TAZ nuclear translocation rate due to the stress fibers or the tensional cytoplasmic F-actin.
n characterizes the weight of how YAP/TAZ activity is dependent on cytoplasmic F-actin. n = 1 is the default
setting that indicates the myosin and cytoplasmic F-actin have a similar effect on regulating YAP/TAZ activity,
and k
ly is the cytoplasmic translocation rate of YAP/TAZ due to active LATS. The mechanism for YAP/TAZ
nuclear import/export is unknown (2), thus we model the YAP/TAZ nuclear transport directly depending on the
stress fibers and LATS as approximation. Importantly, YAP/TAZ nuclear translocation has been shown to
dependent on stress fibers and RhoA in 2D (8). In 3D stress fibers in cells are not as distinct as those present in
the 2D monolayer cultures (59, 60), but still with more F-actin in the cytoplasm in rigid environments in
comparison to the softer ones (61). Thus for cells with no apparent stress fiber structure in 3D, we generally use
the tensional cytoplasmic F-actin filaments, which are characterized by the product of the cytoplasmic F-actin
concentration and active myosin concentration Fcyto(t )myoA(t ), to be an equivalent component of stress
fibers in 3D.
MAL dynamics
MAL in the cytoplasm has been shown to bind with G-actin, which inhibits its ability to go to the nucleus and
associate with SRF to form the activate SRF/MAL complex (50).
(t )
=
dMAL
N
dt
!!
cnm
k
Gactin(t )2
1 + k
mg
(MAL
0
− MAL
N
(t )) − k
ncm
MAL
N
(t ) .
(10)
Here, k
cnm is the rate of MAL translocation from the cytoplasm to the nucleus, k
mg is the decreasing effect due
to cytoplasmic MAL binding with G-actin and retained in the cytoplasm (this inhibition effect is a second-order
hill function), and
ncm is the rate of MAL trans-locating from the nucleus to the cytoplasm.
k
!
Results
We use our computational model to investigate how mechanical signal transmission and YAP/TAZ regulation
depend on pathways of interacting signaling molecules. We study the effect of kinetic inhibition, and perform
local sensitivity analysis to predict the role of crucial model parameters and external variables such as stiffness.
The model not only reproduces the existing experimental results, but also makes a number of new predictions.
First, we study the effect of disturbing adhesion complexes on YAP/TAZ activity represented by the stiffness
response function. Then, we explore the distinction between SRF/MAL and YAP/TAZ sensitivity to ECM
stiffness. More importantly, we investigated the unresolved synergistic effect between the mechanical sensing
and the Hippo pathway of YAP/TAZ through the addition of interaction between LIMK and LATS.
8
Molecular intervention
To demonstrate the fidelity of our model, we reproduced the results observed in experiments with molecular
intervention, so that the model can be readily used to make novel predictions. The YAP/TAZ signaling cascade
in response to mechanical stimuli is studied mostly by molecular interventions targeting the actin cytoskeleton,
such as by adding blebbistatin or latrunculin A (8). We mimicked the inhibition or overexpression process in
our model by upregulating or downregulating the corresponding kinetic parameters. The results are consistent
with the experimental ones from Dupont et al (8) (Figure 2a). Furthermore, we examined the sensitivity of
YAP/TAZ activity to local changes in kinetic parameters (62). It turns out that the parameters relevant to
cell-ECM adhesion are the ones most robustly affecting YAP/TAZ activity (Figure 2b and S2 in the Supporting
Materials), so we study their effect in depth in the following sections.
Figure 2 Molecular intervention of YAP/TAZ activity in stiffness sensing
(a) The results in the computational model regarding molecular interventions are consistent with the experimental results
from previous mechanical regulation studies on YAP/TAZ (8). Here the molecular intervention process is mimicked by
upregulating or downregulating the corresponding kinetic parameters (Supporting Methods in the Supporting Materials).
The soft environment here is 0.5kPa hydrogel for 2D MEC cells, while stiff environment is 20kPa hydrogel(8). Blebbist.:
myosin inhibitor. Y-27632: ROCK inhibitor. C3: RhoA inhibitor. Lat. A: F-actin inhibitor. mDia: over-expression of mDia (a
9
formin that induces F-actin assembly). (b) Local sensitivity analysis of the nondimentionalized model. The local
perturbations of adhesion-relevant parameters ( C , k
and k
df
sf
) are the most robust ones in affecting YAP/TAZ activity
in the model. This is the local sensitivity analysis of YAP/TAZ activity in stiff environment by downregulating the kinetic
parameters individually 10%. The upregulating effect is much weaker than this, since most of the molecular activities are
saturated. The up-regulation in a soft environment also shows that the adhesion is the most robust component in affecting
YAP/TAZ activity (Figure S2 in the Supporting Materials). (c) and (d) are the cell stiffness response function, which is a
YAP/TAZ activity function regulating cellular behavior such as proliferation dependent on external ECM stiffness. (c)
Different amounts of FAK determine the threshold stiffness for cell stiffness sensing by shifting the curves horizontally.
FAK0~1 corresponds to the endogenous FAK amount of NMuMG cell in 3D collagen. Collagen with 25 kPa is a soft
environment and 45 kPa is a stiff environment for 3D NMuMG cells. If FAK is over-expressed to about 3-fold, 20 kPa will
appear as a stiff environment for these cells. (d) The 2D MEC and MSC stiffness sensing identifies 1kPa ECM
environment as soft and 20kPa as stiff, while in 3D NMuMG cells senses 25kPa as soft and 45kPa as stiff. These stiffness
response functions are very alike the ones of manipulating only the FAK total amount of cells in (c).
The initiation of YAP/TAZ signaling converts the biomechanical signals into molecular signals. The first stage
in the cascade, as described in the Methods section and Figure 1, is the phosphorylation level of FAK in
response to ECM stiffness. Overexpressing or downregulating FAK amount has been demonstrated to be able
to rescue or inhibit cell proliferation (52), respectively. Furthermore, overexpression of FAK has shown to
rescue YAP/TAZ activity in serum-starved cells in a PI3K- and LATS-dependent way (63). However, with
sufficient EGF in the environment, we predict that FAK overexpression can rescue YAP/TAZ activity in a soft
ECM environment dependent on the RhoA cascade (Figure 2c). For example, with undisturbed FAK levels, the
cells show high YAP/TAZ activity in ECM with a high stiffness of 45kPa, while they have low YAP/TAZ
activity with 25kPa ECM stiffness. On the other hand, if FAK0 is overexpressed by severalfold, YAP/TAZ
activity is rescued in ECM environment as soft as 20 kPa. More specifically, FAK-overexpression rescues
YAP/TAZ by shifting the stiffness response function leftwards rather than upwards. The transition window in
ECM stiffness between high and low YAP/TAZ activity is also shortened with FAK-overexpression. Thus, we
further predict that in downregulated FAK cells, activating YAP/TAZ activity requires considerably higher
stiffness than in cells with normal FAK levels. That is, the YAP/TAZ activity in downregulated FAK cells can
be rescued by plating them in a much stiffer environment. Changing mDia concentration mainly shifts the
YAP/TAZ stiffness response functions vertically (Figure S3 in the Supporting Materials), and so do many actin
regulators. In addition, increasing FAK concentration results in a narrower range of rigidity sensing and
response in 3D, which displays high sensitivity of YAP/TAZ activity to stiffness between 10 and 15 kPa but
little sensitivity in the range of 25-40 kPa.
In addition, comparing the cell stiffness response functions in 2D and 3D environments (Figure 2d), we found
that their difference in YAP/TAZ activity is very similar to the difference of stiffness response functions with
changing concentrations of FAK (Figure 2c). For example, for normal murine mammary gland cells (NMuMG)
in 3D environments, 45 kPa appears as a stiff environment in terms of YAP/TAZ activity, whereas 25 kPa
appears as soft. On the other hand, for human mammary epithelial cells (MEC), 20kPa appears as stiff whereas
1kPa appears soft. In agreement with the stiffness response function due to FAK concentration change, FAK
has been shown to be more aggregated around the cell membrane in a 2D environment and thus has a higher
total concentration than in a 3D environment, in which FAK is diffused throughout the cytoplasm (24) and has
a lower average concentration. Thus, different localization of FAK in 2D and 3D environments may explain
differences of stiffness sensing and cell behavior in such environments. Due to differences for varying cell
types and dimensions, in the following we talk about the cell behavior and stiffness environment generally,
rather than clarifying specific cell types, dimensions or ECM stiffness values.
10
Comparison between YAP/TAZ and SRF/MAL
In a single-cell mechanical regulation of YAP/TAZ activity is thought to be transmitted mainly though stress
fibers (8), which are a resultant of both F-actin assembly and myosin activity, while SRF/MAL nuclear
translocalization is mainly regulated by the amount of G-actin in the cytoplasm (49, 50). One of the hypotheses
presented in the literature is that due to the resultant interaction of F-actin and myosin, YAP/TAZ nuclear
localization is more sensitive to the stiffness of the ECM than SRF/MAL, which is only affected by actin. To
better study the effect of each molecular perturbation on the YAP/TAZ nuclear translocation in comparison
with MAL nuclear translocation, we define the fold-change of nuclear YAP/TAZ fraction when comparing soft
and stiff ECM environment,
FoldChange =
(YAP / TAZ N )stiff
− (YAP / TAZ N )soft
(YAP / TAZ N )soft
. (11)
This fold-change characterizes the sensitivity of YAP/TAZ activity with regard to stiffness, that is, soft versus
stiff. SRF/MAL sensitivity to the ECM environment is also defined in the same way as Eq. 11. YAP/TAZ
activity was found to be more sensitive to ECM stiffness than SRF/MAL signaling (Figure 3b and S4 in the
Supporting Materials). SRF/MAL is clearly sensitive to F/G-actin, while there is no clear evidence of
SRF/MAL being a genuine mechano-sensor. In addition, YAP/TAZ nuclear fold-change changes more
dramatically by downregulating the activation rates of FAK and RhoA than by upregulating them. It also shows
that changing FAK activation has a larger impact on YAP/TAZ fold change than changing RhoA activation.
The reduction of YAP/TAZ stiffness sensitivity resulting from downregulated FAK (reducing FAK activation
rate) can be rescued by upregulating RhoA (increasing the RhoA activation rate). However, the reduction of
YAP/TAZ stiffness sensitivity by reducing either the mDia or ROCK activation rate cannot be rescued by
increasing the other (Figure S4 in the Supporting Materials). This is due to the fact that mDia and ROCK affect
actin and myosin separately. Hence, the model gives rise to a new approach to compare the sensitivities
quantitatively.
11
Figure 3 YAP/TAZ activity sensitivity analysis and comparison with SRF/MAL
(a) shows YAP/TAZ nuclear localization under different stiffness environment, in comparison with MAL nuclear
translocation. Here MAL nuclear translocation is not as robust as YAP/TAZ in response to stiffness change, likely due to
YAP/TAZ having a myosin-dependent amplification effect other than the common regulator G-actin/F-actin in SRF/MAL
activity. (b) shows fold-change of YAP/TAZ and MAL activities depending on the kinetic parameters, such as the activation
rate of FAK ( k
/ k
sf 0
sf
) and RhoA ( k
/ k
fkρ
fkρ0
), when switching matrix from a soft to a stiff one. In addition, YAP/TAZ
keeps showing higher sensitivity than MAL translocation fold-change in response to different stiffnesses, and the shape of
the activated region reveals that downregulating FAK and RhoA has a more robust effect in regulating the YAP/TAZ
nuclear localization than upregulating it.
The synergistic effect between the Hippo and mechano-sensing pathways
Our previous analyses are carried out with a constant LATS condition. There is a puzzling fact arising from
simultaneous changes in LATS activity and cell tension due to mechanical signals: Dupont et al (8) showed that
LATS kinase activity is independent of cell tension, while Aragona et al (7) showed inhibiting LATS and
capping proteins for F-actin together can promote YAP/TAZ activity to a greater degree than the total
combined effects of adding only siCapZ and adding only siLATS. This is thus a synergy in inhibiting capping
proteins and LATS. In the previous work, this phenomenon is characterized as LATS regulation dominated by
cytoskeleton regulation (7), however, its mechanism is unknown. Our model predicts that this synergistic effect
is from the direct interaction between LATS and LIMK-cofilin. In single-cell system within a soft environment,
though most LATS is inactive, LATS can dramatically reduce the phosphorylation of cofilin and enhances actin
polymerization in an LATS kinase activity-independent way (64). By adding this interaction (the red interaction
12 in Figure 1), our model is able to capture the synergistic effect between the Hippo and mechano-sensing
pathways. Through this interaction, knockout of LATS enhances F-actin assembly. More importantly, it gives
12
rise to the synergistic effect of LATS and CapZ inhibition (Figure 4a). If YAP/TAZ is more dependent on
cytoplasmic F-actin than myosin (n changes from 1 to 2 in Eq. 9), the synergistic effect is larger (Figure 4b).
Figure 4 The synergistic effect between the mechano-sensing and the Hippo pathway.
(a) Adding the interaction between LIMK and LATS, which induces more free active LIMK molecules by adding siLats,
YAP/TAZ is activated synergistically compared with inducing F-actin assembly and adding siLats separately. Condition1:
stiff environment. 2: Soft environment. 3: Soft environment with siLats. 4: Soft environment with siCapZ. 5: Soft
environment with both siLats and siCapZ. CapZ is a capping protein that inhibits F-actin polymerization. Here the effect of
siCapZ on soft environment can be quantified as its corresponding YAP/TAZ activity (H(siCapZ)) abstracting the activity of
YAP/TAZ in soft environment (control: H(soft)). The synergistic effect is quantified as H(siCapZ) – H(soft) + H(siLats) –
H(soft) > H(siCapZ+siLats) – H(soft), i.e., H(siCapZ) + H(siLats) > H(soft) + H(siCapZ+siLats). (b) With a higher
dependence on F-actin than on myosin, the YAP/TAZ showed a more apparent synergistic effect.
Discussion
The key goal of our study is to understand YAP/TAZ activity as a function of mechanical ECM properties. Here
we present a novel mechano-biochemical model that predicts YAP/TAZ activity by integrating adhesion, RhoA
GTPases and actomyosin dynamics. The transduction of the mechanical ECM properties into intracellular
signaling is quantitatively characterized, and we predict how regulating the signaling molecules can make cells
feel different mechanical environments. This gives rise to a direct quantitative function of the mechanical and
intracellular molecule signals, the stiffness response function. We have focused on a few core regulators of
cytoskeleton dynamics, which are sufficient to predict the difference between cytoskeleton-dependent YAP/TAZ
and SRF/MAL activities. Furthermore, the synergistic effect between the LATS and F-actin regulator LIMK
can be fully captured by this model. The model indicates that F-actin may have a larger impact in modulating
YAP activity than myosin. The bulk of the previous studies on cell-matrix interactions have relied on perturbing
signaling molecules in different ECM environment, but mostly characterize only relative upregulating or
downregulating effects. Our study thus provides us with methods to systematically study the sensitivity of
signaling molecules and the environment with the stiffness response function and fold-change function of
different effectors (YAP and SRF). Most importantly, this work provides a quantitative framework for studying
YAP/TAZ activity in response to mechanical sensing and predicts consistent results with the synergistic effect
of mechanical and canonical Hippo signaling.
There are several interesting aspects and predictions of the model that merit further discussion. A key outcome
is that the adhesion is a more robust regulator in affecting YAP/TAZ activities than other cytoskeleton regulators
(Figure 2b). This may be due to two reasons: one is because it is the most upstream signaling molecule that
initiates the signal transmission and others are not as sensitive to the input signals. The other, and perhaps a
more important reason may be that for the RhoA actomyosin cascade, mDia and ROCK are acting
13
complementarily on actin polymerization. A molecule in either of the mDia or ROCK cascades is not as
influential in affecting YAP/TAZ as upstream molecules such as FAK or RhoA. This observations is consistent
with experimental studies focusing on overexpression of mDia (8) where YAP/TAZ activities in both stiff and
soft environment are up-regulated but still have a distinction, due to the additional ROCK cascade. Furthermore,
manipulating FAK will change the range of stiffness sensing where YAP/TAZ activity rises with increasing
stiffness. However, manipulating mDia or other actin regulators will only raise the activity of YAP/TAZ at soft
stiffness, but will not change the range of stiffness sensitivity (Figure S3 in the Supporting Materials). Our
model also shows that the upregulation of mDia or deletion of cofilin upregulates YAP/TAZ through inducing
contractile actin stress fiber, which is consistent with what has been observed previously (65). Secondly, the
inhibition of LATS and CapZ analysis in our model indicates that YAP/TAZ may have a higher dependence on
F-actin than myosin. It remains unclear whether this is due to stress fiber formation itself has a larger
dependence on F-actin, or whether YAP/TAZ is more sensitive to F-actin than myosin. There have been
proposed interactions between myosin and actin cytoskeleton, while whether myosin leads to F-actin assembly
(66, 67) or disassembly (34) is not well studied. Our model in its current form does not include any direct
interactions between myosin and F-actin. The third interesting prediction is that LATS regulates the F-actin
formation via interaction with LIMK/Cofilin, which gives rise to the explanation for the synergistic effect of
YAP/TAZ activity between the mechanosensing and the Hippo pathways. This is consistent with some recent
findings in Drosophila. The overexpression of hpo reduces F-actin at apical sites independent of Yki/YAP, and
partially suppresses the F-actin accumulation caused by knocking down the capping proteins (57). By inhibiting
the binding of LIMK and LATS while preserving the integrity of their individual phosphorylation function, the
synergistic effect of the Hippo and the mechanical transduction will be greatly reduced or diminished. Other
actin regulators, such as ENA (68) and zyxin (69), were found to affect Yki/YAP activity synergistically (70),
most likely through their actin bundles and stress fiber regulation as well. The experimental study of the
YAP/TAZ regulation cascade is difficult, due to the existence of the co-existing alternative interactions and
pathways mentioned above. With our computational model, we are able to investigate the proposed interactions
or possible mechanisms for the YAP/TAZ regulation in a more comprehensive way. This method facilitates an
understanding of the complicated interactions and pathways in an alternative and efficient manner.
There are also some parts in need of further clarifications. Currently the model mostly studies the single cell
stiffness-sensing mechanism in an EGF-sufficient environment, and is not focused on multicellular
environments or cell-cell adhesion or stress. The changing of EGF or serum dose can also be rate-limiting
factors for this stiffness sensing and YAP/TAZ signaling pathway and is not considered in the model. That said,
as mentioned in the methods section, our model can be extended to include other molecules, such as the Hippo
upstream regulators or PI3K, which are also known to affect YAP/TAZ activity. Secondly, it has been
demonstrated that expressing kinase-dead FAK in FAK-/- cells prevented growth, which is consistent with our
model. However, FAK-/- cells exhibit an excessive adhesion formation and uncontrolled proliferation (52).
This suggests that the absence of FAK, rather than inactive FAK, triggers an alternative pathway activation (51,
52, 71). Nevertheless, the FAK-null cells have been shown to be unable to detect soft versus rigid substrata (72).
Thus here our predictions and analysis of FAK activity are restricted to the regime of the downregulation of its
total amount or the phosphorylation rate change, rather than the complete knockout of FAK. There have also
been studies reported in literature showing the mechanical feedback of F-actin stiffening on ECM properties
(73). However, the role of biochemical signaling transduction during this process is not well known and thus
this feedback is not currently included in the model.
Overall, our model provides a platform to study mechanical regulation and molecular crosstalk of YAP/TAZ
upstream regulators. It will be interesting to combine our stiffness-sensing model with models such as the effect
of growth factor stimulation to explore potential interactions between various stimuli. Likewise, it is also
important to extend the model in a multicellular context to capture the in vivo complexity. To this purpose,
14
incorporating cell-cell adhesion and its effect on the dynamics of the canonical Hippo pathway into our model
will provide useful insights. Considering that the mechanical effects can regulate YAP/TAZ independent of the
canonical Hippo pathway, and the cell size control mechanism is very similar to the one of stiffness sensing (7),
we believe that our current model can be useful and quantitatively predictive in a multicellular environment in
elucidating key processes that are too costly to investigate or remain experimentally intractable.
Author Contributions
M.S., F.S. and M.H.Z developed the concepts and designed the models, M.S. implemented the simulations, M.
S. and F. S. performed data analysis, and M.S, F.S. and M.H.Z wrote the article.
Acknowledgements
The authors acknowledge research support from the National Institutes of Health (U01-CA177799) and the
National Science Foundation (DMR-1206635) for this work. We deeply appreciate the input from members of
our lab during this research, and especially Robert J. Seager and Dan Reynolds for useful comments on our
manuscript.
Supporting Citations
References (74–80) appear in the Supporting Material.
Reference
1. Zhao, B., X. Wei, W. Li, R.S. Udan, Q. Yang, J. Kim, J. Xie, T. Ikenoue, J. Yu, L. Li, P. Zheng, K. Ye,
A. Chinnaiyan, G. Halder, Z. Lai, and K. Guan. 2007. Inactivation of YAP oncoprotein by the Hippo
pathway is involved in cell contact inhibition and tissue growth control. Genes Dev. : 2747–2761.
2. Piccolo, S., S. Dupont, and M. Cordenonsi. 2014. The biology of YAP/TAZ: hippo signaling and beyond.
Physiol. Rev. 94: 1287–312.
3. Low, B.C., C.Q. Pan, G. V Shivashankar, A. Bershadsky, M. Sudol, and M. Sheetz. 2014. YAP/TAZ as
mechanosensors and mechanotransducers in regulating organ size and tumor growth. FEBS Lett. 588:
2663–2670.
4. Goulev, Y., J.D. Fauny, B. Gonzalez-Marti, D. Flagiello, J. Silber, and A. Zider. 2008. SCALLOPED
Interacts with YORKIE, the Nuclear Effector of the Hippo Tumor-Suppressor Pathway in Drosophila.
Curr. Biol. 18: 435–441.
5. Ren, F., L. Zhang, and J. Jiang. 2010. Hippo signaling regulates Yorkie nuclear localization and activity
through 14-3-3 dependent and independent mechanisms. Dev. Biol. 337: 303–312.
6. Moore, J.D. 2001. The Ran-GTPase and cell-cycle control. BioEssays. 23: 77–85.
7. Aragona, M., T. Panciera, A. Manfrin, S. Giulitti, F. Michielin, N. Elvassore, S. Dupont, and S. Piccolo.
2013. A mechanical checkpoint controls multicellular growth through YAP/TAZ regulation by
actin-processing factors. Cell. 154: 1047–1059.
8. Dupont, S., L. Morsut, M. Aragona, E. Enzo, S. Giulitti, M. Cordenonsi, F. Zanconato, J. Le Digabel, M.
Forcato, S. Bicciato, N. Elvassore, and S. Piccolo. 2011. Role of YAP/TAZ in mechanotransduction.
Nature. 474: 179–183.
15
9. Discher, D.E., P. Janmey, and Y. Wang. 2005. Tissue Cells Feel and Respond to the Stiffness of Their
Substrate. Sci. . 310 : 1139–1143.
10. Sun, M., A.B. Bloom, and M.H. Zaman. 2015. Rapid Quantification of 3D Collagen Fiber Alignment
and Fiber Intersection Correlations with High Sensitivity. PLoS One. 10: e0131814.
11. Provenzano, P.P., D.R. Inman, K.W. Eliceiri, and P.J. Keely. 2009. Matrix density-induced
mechanoregulation of breast cell phenotype, signaling and gene expression through a FAK-ERK linkage.
Oncogene. 28: 4326–4343.
12. Lo, C.-M., H.-B. Wang, M. Dembo, and Y. Wang. 2000. Cell Movement Is Guided by the Rigidity of
the Substrate. Biophys. J. 79: 144–152.
13. Wu, C., S.B. Asokan, M.E. Berginski, E.M. Haynes, N.E. Sharpless, J.D. Griffith, S.M. Gomez, and J.E.
Bear. 2012. Arp2/3 is critical for lamellipodia and response to extracellular matrix cues but is
dispensable for chemotaxis. Cell. 148: 973–987.
14. Engler, A.J., S. Sen, H.L. Sweeney, and D.E. Discher. 2006. Matrix Elasticity Directs Stem Cell Lineage
Specification. Cell. 126: 677–689.
15. Tang, Y., R.G. Rowe, E.L. Botvinick, A. Kurup, A.J. Putnam, M. Seiki, V.M. Weaver, E.T. Keller, S.
Goldstein, J. Dai, D. Begun, T. Saunders, and S.J. Weiss. 2013. MT1-MMP-Dependent Control of
Skeletal Stem Cell Commitment via a β1-Integrin/YAP/TAZ Signaling Axis. Dev. Cell. 25: 402–416.
16. Berrier, A.L., and K.M. Yamada. 2007. Cell–matrix adhesion. J. Cell. Physiol. 213: 565–573.
17. Harunaga, J.S., and K.M. Yamada. 2011. Cell-matrix adhesions in 3D. Matrix Biol. 30: 363–368.
18. Hanks, S.K., and T.R. Polte. 1997. Signaling through focal adhesion kinase. Bioessays. 19: 137–145.
19. Mitra, S.K., and D.D. Schlaepfer. 2006. Integrin-regulated FAK–Src signaling in normal and cancer cells.
Curr. Opin. Cell Biol. 18: 516–523.
20. Schaller, M.D., J.D. Hildebrand, J.D. Shannon, J.W. Fox, R.R. Vines, and J.T. Parsons. 1994.
Autophosphorylation of the focal adhesion kinase, pp125FAK, directs SH2-dependent binding of
pp60src. Mol. Cell. Biol. . 14 : 1680–1688.
21. Calalb, M.B., T.R. Polte, and S.K. Hanks. 1995. Tyrosine phosphorylation of focal adhesion kinase at
sites in the catalytic domain regulates kinase activity: a role for Src family kinases. Mol. Cell. Biol. 15:
954–963.
22. Paszek, M.J., N. Zahir, K.R. Johnson, J.N. Lakins, G.I. Rozenberg, A. Gefen, C.A. Reinhart-King, S.S.
Margulies, M. Dembo, D. Boettiger, D.A. Hammer, and V.M. Weaver. 2005. Tensional homeostasis and
the malignant phenotype. Cancer Cell. 8: 241–254.
23. Levental, K.R., H. Yu, L. Kass, J.N. Lakins, M. Egeblad, J.T. Erler, S.F.T. Fong, K. Csiszar, A. Giaccia,
W. Weninger, M. Yamauchi, D.L. Gasser, and V.M. Weaver. 2009. Matrix Crosslinking Forces Tumor
Progression by Enhancing Integrin Signaling. Cell. 139: 891–906.
24. Fraley, S.I., Y. Feng, R. Krishnamurthy, D.-H. Kim, A. Celedon, G.D. Longmore, and D. Wirtz. 2010. A
distinctive role for focal adhesion proteins in three-dimensional cell motility. Nat. Cell Biol. 12: 598–
604.
25. Khatiwala, C.B., S.R. Peyton, and A.J. Putnam. 2006. Intrinsic mechanical properties of the extracellular
matrix affect the behavior of pre-osteoblastic MC3T3-E1 cells. Am. J. Physiol. - Cell Physiol. 290:
C1640–C1650.
16
26. Zaman, M.H., L.M. Trapani, A.L. Sieminski, D. MacKellar, H. Gong, R.D. Kamm, A. Wells, D.A.
Lauffenburger, and P. Matsudaira. 2006. Migration of tumor cells in 3D matrices is governed by matrix
stiffness along with cell-matrix adhesion and proteolysis. Proc. Natl. Acad. Sci. 103: 10889–10894.
27. Mak, M., F. Spill, R.D. Kamm, and M.H. Zaman. 2015. Single-Cell Migration in Complex
Microenvironments: Mechanics and Signaling Dynamics. J. Biomech. Eng. .
28. Spill, F., D.S. Reynolds, R.D. Kamm, and M.H. Zaman. 2016. Impact of the physical microenvironment
on tumor progression and metastasis. Curr. Opin. Biotechnol. 40: 41–48.
29. Walcott, S., and S.X. Sun. 2010. A mechanical model of actin stress fiber formation and substrate
elasticity sensing in adherent cells. Proc. Natl. Acad. Sci. . 107 : 7757–7762.
30. Halder, G., S. Dupont, and S. Piccolo. 2012. Transduction of mechanical and cytoskeletal cues by YAP
and TAZ. Nat. Rev. Mol. Cell Biol. 13: 591–600.
31. Harrison, S.M., T. Knifley, M. Chen, and K.L. O Connor. 2013. LPA, HGF, and EGF utilize distinct
combinations of signaling pathways to promote migration and invasion of MDA-MB-231 breast
carcinoma cells. BMC Cancer. 13: 501.
32. Asthagiri, a, C.M. Nelson, a F. Horwitz, and D. a Lauffenburger. 1999. Quantitative relationship
among integrin ligand-binding, adhesion, and signaling via focal adhesion kinase and extracellular
signal-regulated kinase 2. J. Biol. Chem. 274: 27119–27127.
33. Holmes, W.R., B. Lin, A. Levchenko, and L. Edelstein-Keshet. 2012. Modelling cell polarization driven
by synthetic spatially graded Rac activation. PLoS Comput. Biol. 8.
34. Onsum, M., and C. V. Rao. 2007. A mathematical model for neutrophil gradient sensing and polarization.
PLoS Comput. Biol. 3: 0436–0450.
35. Tania, N., J. Condeelis, and L. Edelstein-Keshet. 2013. Modeling the synergy of cofilin and Arp2/3 in
lamellipodial protrusive activity. Biophys. J. 105: 1946–1955.
36. Tania, N., E. Prosk, J. Condeelis, and L. Edelstein-Keshet. 2011. A temporal model of cofilin regulation
and the early peak of actin barbed ends in invasive tumor cells. Biophys. J. 100: 1883–1892.
37. Tao, J., and S.X. Sun. 2015. Active Biochemical Regulation of Cell Volume and a Simple Model of Cell
Tension Response. Biophys. J. 109: 1541–1550.
38. Pollard, T.D., L. Blanchoin, and R.D. Mullins. 2000. Molecular Mechanisms Controlling Actin Filament
Dynamics in Nonmuscle Cells. Annu. Rev. Biophys. Biomol. Struct. 29: 545–576.
39. Sarvestani, A.S. 2013. The effect of substrate rigidity on the assembly of specific bonds at biological
interfaces. Soft Matter. : -.
40. Giannone, G., and M.P. Sheetz. 2006. Substrate rigidity and force define form through tyrosine
phosphatase and kinase pathways. Trends Cell Biol. 16: 213–223.
41. Wozniak, M.A., R. Desai, P.A. Solski, C.J. Der, and P.J. Keely. 2003. ROCK-generated contractility
regulates breast epithelial cell differentiation in response to the physical properties of a
three-dimensional collagen matrix. J. Cell Biol. 163: 583–595.
42. Sakamoto, S., T. Ishizaki, K. Okawa, S. Watanabe, T. Arakawa, N. Watanabe, and S. Narumiya. 2012.
Liprin-controls stress fiber formation by binding to mDia and regulating its membrane localization. J.
Cell Sci. 125: 108–120.
43. Loria, P.M., A. Duke, J.B. Rand, and O. Hobert. 2003. The Mouse Formin mDia1 Is a Potent Actin
17
Nucleation Factor Regulated by Autoinhibition. Curr. Biol. 13: 1317–1323.
44. Narumiya, S., M. Tanji, and T. Ishizaki. 2009. Rho signaling, ROCK and mDia1, in transformation,
metastasis and invasion. Cancer Metastasis Rev. 28: 65–76.
45. Higashida, C., T. Miyoshi, A. Fujita, F. Oceguera-Yanez, J. Monypenny, Y. Andou, S. Narumiya, and N.
Watanabe. 2004. Actin Polymerization-Driven Molecular Movement of mDia1 in Living Cells. Sci. .
303 : 2007–2010.
46. Higashida, C., S. Suetsugu, T. Tsuji, J. Monypenny, S. Narumiya, and N. Watanabe. 2008. G-actin
regulates rapid induction of actin nucleation by mDia1 to restore cellular actin polymers. J. Cell Sci. 121:
3403–3412.
47. Totsukawa, G., Y. Yamakita, S. Yamashiro, D.J. Hartshorne, Y. Sasaki, and F. Matsumura. 2000.
Distinct roles of ROCK (Rho-kinase) and MLCK in spatial regulation of MLC phosphorylation for
assembly of stress fibers and focal adhesions in 3T3 fibroblasts. J. Cell Biol. 150: 797–806.
48. Bamburg, J.R., and B.W. Bernstein. 2010. Roles of ADF/cofilin in actin polymerization and beyond.
F1000 Biol. Rep. 2: 62.
49. Trappmann, B., J.E. Gautrot, J.T. Connelly, D.G.T. Strange, Y. Li, M.L. Oyen, M.A. Cohen Stuart, H.
Boehm, B. Li, V. Vogel, J.P. Spatz, F.M. Watt, and W.T.S. Huck. 2012. Extracellular-matrix tethering
regulates stem-cell fate. Nat. Mater. 11: 642–9.
50. Connelly, J.T., J.E. Gautrot, B. Trappmann, D.W.-M. Tan, G. Donati, W.T.S. Huck, and F.M. Watt.
2010. Actin and serum response factor transduce physical cues from the microenvironment to regulate
epidermal stem cell fate decisions. Nat. Cell Biol. 12: 711–718.
51. Mitra, S.K., D.A. Hanson, and D.D. Schlaepfer. 2005. Focal adhesion kinase: in command and control of
cell motility. Nat Rev Mol Cell Biol. 6: 56–68.
52. Pirone, D.M., W.F. Liu, S.A. Ruiz, L. Gao, S. Raghavan, C.A. Lemmon, L.H. Romer, and C.S. Chen.
2006. An inhibitory role for FAK in regulating proliferation: a link between limited adhesion and
RhoA-ROCK signaling. J. Cell Biol. . 174 : 277–288.
53. Kumar, S., A. Das, and S. Sen. 2014. Extracellular matrix density promotes EMT by weakening cell-cell
adhesions. Mol. Biosyst. 10: 838–50.
54. Zhai, J., H. Lin, Z. Nie, J. Wu, R. Cañete-Soler, W.W. Schlaepfer, and D.D. Schlaepfer. 2003. Direct
Interaction of Focal Adhesion Kinase with p190RhoGEF. J. Biol. Chem. . 278 : 24865–24873.
55. Guilluy, C., V. Swaminathan, R. Garcia-Mata, E.T. O'Brien, R. Superfine, and K. Burridge. 2011. The
Rho GEFs LARG and GEF-H1 regulate the mechanical response to force on integrins. Nat. Cell Biol. 13:
722–727.
56. Bernstein, B.W., and J.R. Bamburg. 2010. ADF/Cofilin: a functional node in cell biology. Trends Cell
Biol. 20: 187–195.
57. Fernández, B.G., P. Gaspar, C. Brás-Pereira, B. Jezowska, S.R. Rebelo, and F. Janody. 2011.
Actin-Capping Protein and the Hippo pathway regulate F-actin and tissue growth in Drosophila.
Development. 138: 2337–2346.
58. Kannan, N., and V.W. Tang. 2015. Synaptopodin couples epithelial contractility to α-actinin-4–
dependent junction maturation. J. Cell Biol. . 211 : 407–434.
59. Toh, Y.-C., C. Zhang, J. Zhang, Y.M. Khong, S. Chang, V.D. Samper, D. van Noort, D.W. Hutmacher,
and H. Yu. 2007. A novel 3D mammalian cell perfusion-culture system in microfluidic channels. Lab
18
Chip. 7: 302–9.
60. LI, S., J. LAO, B.P.C. CHEN, Y.-S. LI, Y. ZHAO, J. CHU, K.-D. CHEN, T.-C. TSOU, K. PECK, and
S.H.U. CHIEN. 2003. Genomic analysis of smooth muscle cells in 3-dimensional collagen matrix.
FASEB J. . 17 : 97–99.
61. Peyton, S.R., P.D. Kim, C.M. Ghajar, D. Seliktar, and A.J. Putnam. 2008. The effects of matrix stiffness
and RhoA on the phenotypic plasticity of smooth muscle cells in a 3-D biosynthetic hydrogel system.
Biomaterials. 29: 2597–2607.
62. Zi, Z. 2011. Sensitivity analysis approaches applied to systems biology models. IET Syst. Biol. 5: 336.
63. Kim, N.-G., and B.M. Gumbiner. 2015. Adhesion to fibronectin regulates Hippo signaling via the FAK–
Src–PI3K pathway. J. Cell Biol. . 210 : 503–515.
64. Yang, X., K. Yu, Y. Hao, D. Li, R. Stewart, K.L. Insogna, and T. Xu. 2004. LATS1 tumour suppressor
affects cytokinesis by inhibiting LIMK1. Nat. Cell Biol. 6: 609–617.
65. Kanellos, G., J. Zhou, H. Patel, R.A. Ridgway, D. Huels, C.B. Gurniak, E. Sandilands, N.O. Carragher,
O.J. Sansom, W. Witke, V.G. Brunton, and M.C. Frame. 2015. ADF and Cofilin1 Control Actin Stress
Fibers, Nuclear Integrity, and Cell Survival. Cell Rep. .
66. Lechler, T., A. Shevchenko, A. Shevchenko, and R. Li. 2000. Direct involvement of yeast type I myosins
in Cdc42-dependent actin polymerization. J. Cell Biol. 148: 363–373.
67. Huckaba, T.M., T. Lipkin, and L. a. Pon. 2006. Roles of type II myosin and a tropomyosin isoform in
retrograde actin flow in budding yeast. J. Cell Biol. 175: 957–969.
68. Krause, M., E.W. Dent, J.E. Bear, J.J. Loureiro, and F.B. Gertler. 2003. Ena/VASP proteins: regulators
of the actin cytoskeleton and cell migration. Annu. Rev. Cell Dev. Biol. 19: 541–564.
69. Harborth, J., S.M. Elbashir, K. Bechert, T. Tuschl, and K. Weber. 2001. Identification of essential genes
in cultured mammalian cells using small interfering RNAs. J. Cell Sci. 114: 4557–4565.
70. Gaspar, P., M. V. Holder, B.L. Aerne, F. Janody, and N. Tapon. 2015. Zyxin antagonizes the FERM
protein expanded to couple f-actin and yorkie-dependent organ growth. Curr. Biol. 25: 679–689.
71. Playford, M.P., K. Vadali, X. Cai, K. Burridge, and M.D. Schaller. 2008. Focal Adhesion Kinase
regulates cell–cell contact formation in epithelial cells via modulation of Rho. Exp. Cell Res. 314: 3187–
3197.
72. Wang, H.-B., M. Dembo, S.K. Hanks, and Y. Wang. 2001. Focal adhesion kinase is involved in
mechanosensing during fibroblast migration. Proc. Natl. Acad. Sci. . 98 : 11295–11300.
73. Provenzano, P.P., K.W. Eliceiri, J.M. Campbell, D.R. Inman, J.G. White, and P.J. Keely. 2006. Collagen
reorganization at the tumor-stromal interface facilitates local invasion. BMC Med. 4: 38.
74. Yarmola, E.G., T. Somasundaram, T. a. Boring, I. Spector, and M.R. Bubb. 2000. Actin-latrunculin a
structure and function: Differential modulation of actin-binding protein function by latrunculin A. J. Biol.
Chem. 275: 28120–28127.
75. Dawes, A.T., and L. Edelstein-Keshet. 2007. Phosphoinositides and Rho proteins spatially regulate actin
polymerization to initiate and maintain directed movement in a one-dimensional model of a motile cell.
Biophys. J. 92: 744–768.
76. Feng, J., M. Ito, Y. Kureishi, K. Ichikawa, M. Amano, N. Isaka, K. Okawa, A. Iwamatsu, K. Kaibuchi,
D.J. Hartshorne, and T. Nakano. 1999. Rho-associated Kinase of Chicken Gizzard Smooth Muscle. J.
19
Biol. Chem. . 274 : 3744–3752.
77. Ji, H., H. Tang, H. Lin, J. Mao, L. Gao, J.I.A. Liu, and T. Wu. 2014. Rho / Rock cross ‑ talks with
transforming growth factor ‑ β / Smad pathway participates in lung fibroblast ‑ myofibroblast
differentiation. Biomed. Reports. : 787–792.
78. Cirit, M., M. Krajcovic, C.K. Choi, E.S. Welf, A.F. Horwitz, and J.M. Haugh. 2010. Stochastic model of
integrin-mediated signaling and adhesion dynamics at the leading edges of migrating cells. PLoS
Comput. Biol. 6.
79. Song, X., X. Chen, H. Yamaguchi, G. Mouneimne, J.S. Condeelis, and R.J. Eddy. 2006. Initiation of
cofilin activity in response to EGF is uncoupled from cofilin phosphorylation and dephosphorylation in
carcinoma cells. J. Cell Sci. . 119 : 2871–2881.
80. Pollard, T. 1986. Rate constants for the reactions of ATP-and ADP-actin with the ends of actin filaments.
J. Cell Biol. 103: 2747.
20
Supporting Materials
Supporting Methods
Molecular intervention
When latrunculin A is employed in vitro, the depolymerization of actin filaments
results from sequestration of monomeric actin by latrunculin (1), and the effect in our
model is mimicked by changing the polymerization rate of F-actin in the cytoplasm to
1/10 of the original. Overexpression of mDia is mimicked by changing mDia0 to 5-
fold. Blebbistatin and Y-27632 inhibit myosin and ROCK, thus we mimic their
effects by changing the deactivation rates of myosin
kdmy
and the one of ROCK to 10-
fold.
Local sensitivity analysis of YAP/TAZ activity on kinetic parameters
Each kinetic parameter is locally up- or down-regulated by 10% individually, the
YAP/TAZ activity, quantified by YAP/TAZ nuclear translocation fraction, is
calculated
ParameterSensitivity =
!!
YAP /TAZ
(k + Δk) − YAP /TAZ
N
YAP /TAZ
(k)
N
(k)
N
/(
Δk
k
) .
Here Δk / k = 10% . If parameter sensitivity is 1, the dependence on the correspondent
kinetic parameters is linear. If it is larger than 1, it shows the YAP/TAZ activity is
highly dependent on the parameters in a way that is more dramatic than the linear
dependence, while if it is smaller than 1, then it is the other way around (2).
Supporting Figures
Figure S1 The FAK activity model comparison.
Our model adapts from the previous Michaelis-Menten model of FAK activation rate
dependent on ECM ligand density (3, 4). After incorporating the ECM stiffness in the model,
we found that the second-order hill function matches the experimental data from previous
work (5), with the Michaelis-Menten model not able to capture the increase of FAK activity
from soft to stiff environments.
1
Figure S2 Local sensitivity analysis of YAP/TAZ model in soft environment
This shows the local sensitivity in YAP/TAZ model by up-regulating the kinetic parameters in
a soft environment. In consistent with Figure 2b, the adhesion components in the model have
the largest impact on the YAP/TAZ activity.
Figure S3 stiffness response function with changing protein contents
(a) and (b) are active RhoA fraction and stress fiber content changing along with the stiffness
under different amount of FAK. The RhoA activity goes plateau as the stiffness increases.
Both the sensitivity range and the plateau position of RhoA change with FAK intervention. (c)
The YAP/TAZ stiffness response function shifts vertically by changing total amount of mDia
for 3D NMuMG cells as an example. mDia0~1 corresponds the endogenous mDia amount of
NMuMG cell in 3D collagen. (d) The stress fiber content stiffness response function changes
along the total amount of mDia. The stress fiber content change with stiffness gives rise to the
2
YAP/TAZ stiffness response. As predicted in YAP/TAZ activity, the FAK shifts the stress fiber
content stiffness response function horizontally, while other actin molecules such as mDia
shifts the stress fiber content stiffness response function vertically.
Figure S4 The kinetic parameter analysis of ROCK and mDia
Using the same method as in Figure 3b, we analyze the activation rate of ROCK and mDia on
YAP/TAZ sensitivity to ECM stiffness, and also compare with SRF/MAL sensitivity. Similar
with regulating adhesion and RhoA components, down-regulating ROCK and mDia shows a
dramatic effect in the YAP/TAZ sensitivity to stiffness. However, SRF/MAL shows a relative
monotonic dependence on mDia activation, which regulates F-actin polymerization. This
indicates that SRF/MAL increase rate dependence on F-actin is also monotonic and the
increase rate of SRF/MAL in stiff environment is larger than the one in soft environment. The
YAP/TAZ sensitivity to stiffness pattern shows that by down-regulating the ROCK and mDia,
the YAP/TAZ decrease rate in soft environment is first lower than the one in stiff environment,
and then rises the other way. Comparing with the dependence on FAK and RhoA sensitivity
analysis, the region of high YAP/TAZ activity is smaller, indicating that there are other more
robust regulators limiting the active regime of YAP/TAZ activity.
Supporting nondimentionalized model
The nondimentionalization has been carried out for the model. The refined parameters
are in the Table S1. And we mainly focus on the ECM stiffness in our paper and thus
treat ligand density constant. The model is applicable with changing ligand density, as
it is adapted from the ligand density dependent adhesion model (3, 4). The variables
of proteins are normalized per cell. Most of the proteins involved in YAP/TAZ
stiffness sensing have been found to be constant under stiff and soft environment, for
example, the total amount of FAK (2), RhoA (2), total actin (3), and YAP/TAZ (4).
dfak(t )
dt
= ksf
E
mol(t )2
2
C
+ E
mol(t )2
(1 − fak(t )) − kdf fak(t ) (S1),
3
drhoA(t )
!!
dt
= k
fkρ
(γ( fak(t ))2 + 1)(1 − rhoA(t ))− k
rhoA(t ) (S2),
dρ
drock(t )
= k
dt
drock(t )
= k
dt
rhoA(t )(1 − rock(t )) − k
rρ
rock(t ) (S3),
drock
rhoA(t )(1 − rock(t )) − k
rρ
rock(t ) (S4),
drock
dmyo(t )
!!
dt
= k
mr
(εT(rock(t )) + 1)(1 − myo(t )) − k
dmy
myo(t ) (S5),
dlimk(t )
dt
= k
lr(τT(rock(t )) + 1)(1 − limk(t )) − k
dl
limk(t ) (S6),
dcofilin(t )
dt
= k
turn−over
(1 − cofilin(t )) − k
(1 − k
lats
0
ll
cr
)limk 2(t )cofilin(t ) (S7),
dfcyto(t )
dt
= k
ra(αT(mdia(t )) + 1)(1 − factin(t )) − (kdep
+ k fc1cofilin(t )) fcyto(t ) (S8),
= (k
CN
+ kCY fcyton(t )myoA(t ))(1 − yaptaz(t )) − (k
NC
+ klylats(t )) yaptaz(t )
dyaptaz(t )
dt
(S9),
dmal(t )
=
dt
k
cnm
(1 − factin(t ))2
(1 − mal(t )) − k
mal(t ) (S10),
ncm
Table S1. The parameter values used in the nondimentionalized model.
Parameter
k
!
k
sf
df
C
C
fkρ
k
!
γ
k
k
dρ
rρ
k
drock
k
mρ
Value
0.0175 s-1
0.035 s-1
45 kPa for 3D
NMuMG cells
25 kPa for 2D MEC
and MSC
0.018 s-1
500
0.625 s-1
2.2 s-1
0.8 s-1
1 s-1
Reference
Kumar et al (4)
Kumar et al (4)
Estimated from Provenzano et al (5)
Estimated from Dupont (6)
Estimated from Provenzano et al (5)
Estimated from Provenzano et al (5)
Sako et al (7)
Edelstein-Keshet et al (8, 9)
Estimated from Feng et al (10) and Ji et
al (11)
Estimated
4
k
dmDia
k
mr
ε
k
dmy
!
k
lr
τ
k
dl
!
k
turn−over
!
k
cr
k
ll
k
ra
!
α
k
dep
!
k
fc1
k
CN
k
CY
k
NC
k
ly
k
cnm
1 s-1
0.015 s-1
40
0.067s-1
0.07 s-1
200
2 s-1
0.04 s-1
0.7 s-1
0.8 s-1
0.4 s-1
40
3.5 s-1
8
0.1 s-1
20 s-1
3 s-1
6 s-1
0.2s-1
Estimated
Estimated
Estimated
Cirit et al (12)
Estimated from Tania et al and Song et
al (13, 14)
Estimated
Estimated from Tania et al and Song et
al (13, 14)
Tania et al (13)
Tania et al (13)
Estimated from Yang et al (15)
Estimated
Estimated from Higashida et al (16)
Bonder et al (17)
Estimated from Tania et al and Song et
al (13, 14)
Estimated
Estimated
Estimated
Estimated
Estimated from Connelly et al (18)
Table S2. Redefinition of the parameters in the nondimentionalized model
Parameters
in the
nondiment-
ionalized
model
C
γ
k
rρ
k
mρ
ε
τ
Parameters in the
dimentionalized model
C / LD
!!
γFAK
2
0
RhoA
k
!!
rρ
k
!!
mρ
RhoA
0
0
εROCK
!!
τROCK
!!
0
0
Parameters
in the
nondiment-
ionalized
model
k
!
cr
k
ll
α
k
k
fc1
CY
k
ly
Parameters in the
dimentionalized model
LIMK
k
cr
2
0
LATS
k
!!
ll
0
αmDia
!!
0
fc1Cofilin0
nMyo0
actin0
k
!!
k
!!
CY
k
ly
!!
LATS0
5
k
ra
fak
!rhoA
rock
FAK
k
ra(1 + β)
P / FAK
0
GTP / RhoA
RhoA
0
ROCK
A / ROCK
0
k
cnm
!limk
cofilin
fcyto
yaptaz
k
!!
cnm
/k
mg
(1 + β)2
LIMK
A / LIMK
0
/Cofilin
0
Cofilin
!!
NP
Fcyto /actin
!!
0
0
0
mal
mDia
Myo
!!
A
/ Myo
A / mDia
(YAP /TAZ )N /(YAP /TAZ )
!!
mdia
!myo
Supporting Reference
1. Yarmola, E.G., T. Somasundaram, T. a. Boring, I. Spector, and M.R. Bubb.
2000. Actin-latrunculin a structure and function: Differential modulation of
actin-binding protein function by latrunculin A. J. Biol. Chem. 275: 28120–
28127.
Zi, Z. 2011. Sensitivity analysis approaches applied to systems biology models.
IET Syst. Biol. 5: 336.
N / MAL
MAL
!!
2.
0
0
3. Asthagiri, a, C.M. Nelson, a F. Horwitz, and D. a Lauffenburger. 1999.
Quantitative relationship among integrin ligand-binding, adhesion, and
signaling via focal adhesion kinase and extracellular signal-regulated kinase 2.
J. Biol. Chem. 274: 27119–27127.
4. Kumar, S., A. Das, and S. Sen. 2014. Extracellular matrix density promotes
5.
EMT by weakening cell-cell adhesions. Mol. Biosyst. 10: 838–50.
Provenzano, P.P., D.R. Inman, K.W. Eliceiri, and P.J. Keely. 2009. Matrix
density-induced mechanoregulation of breast cell phenotype, signaling and
gene expression through a FAK-ERK linkage. Oncogene. 28: 4326–4343.
6. Dupont, S. Role of YAP/TAZ in cell-matrix adhesion-mediated signalling and
7.
mechanotransduction. Exp. Cell Res. .
Sako, Y., K. Hibino, T. Miyauchi, Y. Miyamoto, M. Ueda, and T. Yanagida.
2000. Single-Molecule Imaging of Signaling Molecules in Living Cells. Single
Mol. 1: 159–163.
8. Holmes, W.R., B. Lin, A. Levchenko, and L. Edelstein-Keshet. 2012.
Modelling cell polarization driven by synthetic spatially graded Rac activation.
PLoS Comput. Biol. 8.
9. Dawes, A.T., and L. Edelstein-Keshet. 2007. Phosphoinositides and Rho
proteins spatially regulate actin polymerization to initiate and maintain directed
movement in a one-dimensional model of a motile cell. Biophys. J. 92: 744–
768.
10. Feng, J., M. Ito, Y. Kureishi, K. Ichikawa, M. Amano, N. Isaka, K. Okawa, A.
Iwamatsu, K. Kaibuchi, D.J. Hartshorne, and T. Nakano. 1999. Rho-associated
Kinase of Chicken Gizzard Smooth Muscle. J. Biol. Chem. . 274 : 3744–3752.
Ji, H., H. Tang, H. Lin, J. Mao, L. Gao, J.I.A. Liu, and T. Wu. 2014. Rho /
Rock cross ‑ talks with transforming growth factor ‑ β / Smad pathway
participates in lung fibroblast ‑ myofibroblast differentiation. Biomed. Reports. :
787–792.
11.
6
12. Cirit, M., M. Krajcovic, C.K. Choi, E.S. Welf, A.F. Horwitz, and J.M. Haugh.
2010. Stochastic model of integrin-mediated signaling and adhesion dynamics
at the leading edges of migrating cells. PLoS Comput. Biol. 6.
13. Tania, N., E. Prosk, J. Condeelis, and L. Edelstein-Keshet. 2011. A temporal
model of cofilin regulation and the early peak of actin barbed ends in invasive
tumor cells. Biophys. J. 100: 1883–1892.
14. Song, X., X. Chen, H. Yamaguchi, G. Mouneimne, J.S. Condeelis, and R.J.
Eddy. 2006. Initiation of cofilin activity in response to EGF is uncoupled from
cofilin phosphorylation and dephosphorylation in carcinoma cells. J. Cell Sci. .
119 : 2871–2881.
15. Yang, X., K. Yu, Y. Hao, D. Li, R. Stewart, K.L. Insogna, and T. Xu. 2004.
LATS1 tumour suppressor affects cytokinesis by inhibiting LIMK1. Nat. Cell
Biol. 6: 609–617.
16. Higashida, C., S. Suetsugu, T. Tsuji, J. Monypenny, S. Narumiya, and N.
Watanabe. 2008. G-actin regulates rapid induction of actin nucleation by
mDia1 to restore cellular actin polymers. J. Cell Sci. 121: 3403–3412.
17. Pollard, T. 1986. Rate constants for the reactions of ATP-and ADP-actin with
the ends of actin filaments. J. Cell Biol. 103: 2747.
18. Connelly, J.T., J.E. Gautrot, B. Trappmann, D.W.-M. Tan, G. Donati, W.T.S.
Huck, and F.M. Watt. 2010. Actin and serum response factor transduce
physical cues from the microenvironment to regulate epidermal stem cell fate
decisions. Nat. Cell Biol. 12: 711–718.
7
|
1302.3196 | 1 | 1302 | 2013-02-13T19:28:03 | Biophysics of filament length regulation by molecular motors | [
"physics.bio-ph",
"q-bio.SC"
] | Regulating physical size is an essential problem that biological organisms must solve from the subcellular to the organismal scales, but it is not well understood what physical principles and mechanisms organisms use to sense and regulate their size. Any biophysical size-regulation scheme operates in a noisy environment and must be robust to other cellular dynamics and fluctuations. This work develops theory of filament length regulation inspired by recent experiments on kinesin-8 motor proteins, which move with directional bias on microtubule filaments and alter microtubule dynamics. Purified kinesin-8 motors can depolymerize chemically-stabilized microtubules. In the length-dependent depolymerization model, the rate of depolymerization tends to increase with filament length, because long filaments accumulate more motors at their tips and therefore shorten more quickly. When balanced with a constant filament growth rate, this mechanism can lead to a fixed polymer length. However, the mechanism by which kinesin-8 motors affect the length of dynamic microtubules in cells is less clear. We study the more biologically realistic problem of microtubule dynamic instability modulated by a motor-dependent increase in the filament catastrophe frequency. This leads to a significant decrease in the mean filament length and a narrowing of the filament length distribution. The results improve our understanding of the biophysics of length regulation in cells. | physics.bio-ph | physics |
Biophysics of filament length regulation by molecular motors
Hui-Shun Kuan
Program in Chemical Physics and Biofrontiers Institute, University of Colorado at Boulder
M. D. Betterton
Department of Physics and Biofrontiers Institute, University of Colorado at Boulder∗
(Dated: September 16, 2018)
Regulating physical size is an essential problem that biological organisms must solve from
the subcellular to the organismal scales, but it is not well understood what physical prin-
ciples and mechanisms organisms use to sense and regulate their size. Any biophysical
size-regulation scheme operates in a noisy environment2 and must be robust to other cel-
lular dynamics and fluctuations. This work develops theory of filament length regulation
inspired by recent experiments on kinesin-8 motor proteins, which move with directional
bias on microtubule filaments and alter microtubule dynamics. Purified kinesin-8 motors
can depolymerize chemically-stabilized microtubules. In the length-dependent depolymer-
ization model, the rate of depolymerization tends to increase with filament length, because
long filaments accumulate more motors at their tips and therefore shorten more quickly.
When balanced with a constant filament growth rate, this mechanism can lead to a fixed
polymer length. However, the mechanism by which kinesin-8 motors affect the length of
dynamic microtubules in cells is less clear. We study the more biologically realistic problem
of microtubule dynamic instability modulated by a motor-dependent increase in the filament
catastrophe frequency. This leads to a significant decrease in the mean filament length and
a narrowing of the filament length distribution. The results improve our understanding of
the biophysics of length regulation in cells.
I.
INTRODUCTION
A fundamental question in biology is how organisms control the size of subcellular structures,
cells, organs, and whole organisms. The physical principles underlying the sensing and control of
size in biology are not well understood; indeed, whether there are general principles or mechanisms
in size control is unclear [1 -- 3]. In particular, the regulation of polymer length is important for
the organization of the cellular cytoskeleton. Regulation of cytoskeletal filaments affects both the
size of subcellular organelles such as the mitotic spindle [4, 5] and the structure of cells themselves
[6, 7]. Microtubules are an important cytoskeletal filament that contribute to cell structure, affect
the distribution of other cytoskeletal filaments, move chromosomes during cell division, and serve
as tracks for transport within the cell. This paper focuses on the regulation of microtubule length.
Microtubules undergo complex, nonequilibrium polymerization dynamics, known as dynamic
instability, characterized by stochastic switching between distinct growing and shrinking states.
When dynamic microtubules polymerize in vitro from purified tubulin protein, dynamic instability
leads to a broad distribution of polymer lengths [8]. However, in cells other proteins are also
present which modify microtubule dynamics, particularly at the plus ends of microtubules [9].
This allows cells to control tubulin polymerization dynamics to give microtubules of regulated
length. A relatively well-studied example of length regulation of microtubule-based structures is
the control of flagellar length in Chlamydomonas reinhardtii, where the assembly and disassembly
∗ [email protected]
2
of the flagellum is controlled to give a fixed flagellar length [10]. However, in general it is not well
understood how cells control the length of their microtubules.
Recently an example of physically-based microtubule length detection and control was proposed,
based on motor proteins that walk with directional bias on a microtubule and shorten a stabilized
(non-polymerizing) microtubule from its plus end [11 -- 13]. If the motors are processive (unbinding
slowly from the microtubule), then a long filament can accumulate large numbers of motors at
its end, and the shortening rate is high; for a short filament fewer motors reach the end, and
shortening slows. This length-dependent depolymerization has been demonstrated for the kinesin-8
protein Kip3p moving on stabilized microtubules. The physical interactions of motors moving on
the filament allow a physical process (the rate of depolymerization) to vary with the filament length,
thereby allowing sensing of the length [12 -- 14]. By coupling this length-sensitive depolymerization
with other processes (for example, a constant filament growth rate) a specific filament length or
narrow filament-length distribution could in principle be achieved [12].
To understand the biological relevance of length-dependent depolymerization, it is important
to make a connection between the biophysically measured effects of purified proteins on stabilized
microtubules and the more complex situation in cells. Stabilized microtubules have little or no
intrinsic length dynamics, while in cells microtubules undergo dynamic instability. Other proteins
can also modify microtubule dynamics. Therefore, the kinesin-8 length-dependent depolymeriza-
tion process will be affected by microtubule length fluctuations and the presence of other proteins
at microtubule tips. In general, any biophysical mechanism of length regulation must be robust to
noise in the cellular environment.
Recent work suggests that direct length-dependent depolymerization may not be occurring in
cells; instead, kinesin-8 proteins may act to promote catastrophe (the transition from growing to
shrinking). Extensive experiments have demonstrated that deletions or knockdowns of kinesin-
8 proteins in cells result in longer microtubules and mitotic spindles as well as an increase in
chromosome loss in mitosis [11, 15 -- 21]. Other work has shown that kinesin-8 activity is associated
with destabilization of microtubules and other alterations in microtubule dynamics [11 -- 13, 15, 20,
22 -- 35]. While the molecular mechanisms of kinesin-8 protein function are not clear, it appears
that not all kinesin-8 proteins are able to depolymerize stabilized microtubules [26, 35, 36]. Both
experimental evidence [11, 25, 29, 35] and modeling work [37] suggest that in cells kinesin-8 proteins
may act to promote catastrophe of dynamic microtubules. Therefore, it is necessary to understand
the consequences of length-dependent changes in microtubule dynamic instability to predict the
effects of these motors in cells. This will improve our general question of how length-sensing
mechanisms are altered by fluctuations and dynamics in biological systems.
Previous theory and modeling work has addressed aspects of kinesin 8 behavior and length
regulation. Several papers have focused on modeling the physical effects important to describe
the length-dependent depolymerization of otherwise static filaments [13, 14, 38] or filaments with
simplified polymerization kinetics [39 -- 41]. To our knowledge, previous work has not examined
the effect of catastrophe-promoting motors on the length distribution of microtubules undergoing
dynamic instability. Tischer et al. used a similar formalism to that in this paper in a model for
how length-dependent microtubule catastrophe and rescue rates affect the density of cargo-carrying
motors along microtubules, an effect that could be used to target cargo delivery to specific cellular
locations [42].
In this paper, we develop a simplified physical theory to compare two scenarios for length regula-
tion: for length regulation by depolymerization we calculate the steady-state length that is reached
by a constantly growing filament balanced by depolymerizing motors, while for length regulation by
altering catastrophe we consider filaments undergoing dynamic instability with alterations in the
dynamics due to motors. We consider two possible mechanisms of motor action at the microtubule
tip, both the minimal model in which motor effects (depolymerization or catastrophe) increase
3
FIG. 1. Schematic of the model. (A) Model of length regulation by depolymerization. Filament growth at
the plus end is balanced by motor-induced depolymerization. Motors bind to and unbind from the filament,
move toward the filament plus end, and catalyze removal of filament subunits from the plus end. This leads
to a length-dependent depolymerization rate, so a single filament length is favored, depending on the model
parameters.
(B) Model of length regulation by altering catastrophe. The filament undergoes dynamic
instability at its plus end, characterized by stochastic transitions between growing (blue) and shrinking
(green) states. Motors bind to and unbind from the filament, move toward the filament plus end, and
catalyze catastrophe (transition from the growing to the shrinking state) at the filament plus end. Motor
effects make the catastrophe frequency length dependent, which leads to a broad distribution of filament
lengths determined by the properties of the motors.
in proportion to the motor density [14, 38, 41] and the cooperative model in which motor effects
(depolymerization or catastrophe) increase in proportion to the flux of motors to the filament end
[13]. These two models show important differences in their effects on length regulation, suggesting
that cellular length regulation could be sensitive to the precise mechanism. We find consistent
qualitative agreement between mean-field theory and stochastic simulation; in some parameter
regimes the two approaches agree quantitatively.
II. MOTOR DYNAMICS ALONG FILAMENT
The mean-field density of the motors along the filament, ρ(x, t), in units of motors per unit
length is described by [43]
∂ρ
∂t
= −v
∂
∂x
(cid:18)
(cid:20)
ρ
(cid:19)(cid:21)
(cid:18)
(cid:19)
1 − ρ
ρmax
+ konc
1 − ρ
ρmax
− koff ρ.
(1)
On the right hand side, the first term describes biased motion of the motors with speed v, where
crowding effects reduce the motor flux [43] and ρmax is the maximum possible motor density.
The second term describes binding of motors to unoccupied sites at rate per unit length konc.
The third term describes unbinding of motors from occupied sites at rate koff . This formulation
assumes a continuum limit in which the lattice spacing a → 0 so that the motor density can be
treated as continuous in x. The bulk motor concentration c is assumed constant, i.e., the binding
of motors to the filament is assumed not to deplete the pool of motors in the bulk. Note that we
do not consider protofilament interactions within a microtubule, so we are effectively considering
a single-protofilament microtubule (figure 1).
For relatively low motor density, we neglect crowding effects in the drift term, which makes
the density equation linear. In addition, we work with the motor fractional occupancy p(x, t) =
(A) Length regulation by depolymerization(B) Length regulation by altering catastropheGrowing filamentShrinking filamentMotor-inducedcatastropheGrowing filamentMotor-induceddepolymerizationρ(x, t)/ρmax, so the density equation can be written
∂p
∂t
= −v
∂p
∂x
+
konc
ρmax
(1 − p) − koff p.
4
(2)
With the initial condition p(x, t) = 0 at time t = 0 and the boundary condition p(x = 0, t) = 0,
the solution to this equation is
(cid:40)
p(x, t) =
p0(1 − e−t/τ ),
x ≥ vt (short time)
p0(1 − e−x/λ), x < vt (long time)
(3)
The steady-state occupancy away from the filament ends is p0 = konc/(koff ρmax + konc), the time
scale τ = 1/(koff + konc/ρmax), and the length scale λ = v/(koff + konc/ρmax) = vτ . At long time,
the density approaches the steady state profile which is constant away from the filament end but
has a boundary layer for small x where transport effects and boundary conditions change the motor
density away from p0 [39, 43, 44].
III. LENGTH REGULATION BY DEPOLYMERIZATION
Here we study the regulation of filament length assuming the motors directly depolymerize
the filament from its plus end, an effect which is balanced by a constant rate of filament growth
(figure 1A). This approach neglects fluctuations due to microtubule dynamic instability, and so
the resulting length is deterministically reached. We suppose that a filament grows with constant
speed u.
We consider two simple models for the dynamics of the plus end end. (We assume that the
filament minus end is not dynamic.) For density-controlled depolymerization, the motor-
induced depolymerization rate is proportional to the motor occupancy at the end [14, 38, 41]. For
flux-controlled depolymerization, the motor-induced depolymerization rate is proportional to
the motor flux to the end [13, 38]. We assume that the motors move faster than the growth
(v > u), so the motors track the end as observed experimentally. We therefore assume that the
motor occupancy away from the filament end reaches the steady-state value p(x) = p0(1 − e−x/λ).
Dimensional analysis suggests that the filament length reached should be related the boundary
layer length scale λ = v/(koff + konc/ρmax) = vτ , the obvious length scale that can be constructed
from the rates in the problem. However, the steady-state filament length is quite different from λ
and is controlled by the dynamics at the end of the filament.
This model is related to recent work that also considered the balance between depolymerizing
motors and filament kinetics described by constant growth [39, 41] or treadmilling [40]. The model
here is simplified compared to the previous work, to allow the derivation of analytic expressions
for the length achieved and to allow comparison to the results for filaments undergoing dynamic
instability.
A. Density-controlled depolymerization
In the density-controlled model, we assume that the depolymerization rate is proportional to
the motor density at the terminal site of the microtubule [14]. We assume that only the motor
occupancy at the last site of the filament is important for depolymerization, i.e., we neglect possible
cooperative effects. Define pe(t) to be the average motor occupancy at the last site on the filament,
5
FIG. 2. Filament dynamics and steady-state filament length for length regulation by depolymerization. Blue
(top), density-controlled depolymerization; red (bottom), flux-controlled depolymerization. Left, example
trace of filament length versus time in the stochastic simulation for c = 10 nM. Middle, normalized filament
length distribution in the stochastic simulation for c = 10 nM, averaged from 10 stochastic simulations after
removal of the initial transient. Right, steady-state filament length predictions of mean-field theory and the
stochastic simulation (error bars are standard deviations of steady-state length distributions). The steady-
state length varies rapidly with the bulk motor concentration. The two models give similar predictions, and
the approximate expressions (eqns 7 and 10) for the steady-state length are within a factor of two of the
exact expressions (eqns 6 and 9) except for very low bulk motor concentrations. The mean-field theory uses
the parameters v = 3 µm min−1, kon = 2 nM−1µm−1 min−1, koff = 0.25 min−1, kend
off = 1.45 min−1,
w = 1.025 µm min−1, a = 8 nm, δ = 8 nm, and ρmax = 125 µm−1; for the density-controlled model u = 1.0
µm min−1 while for the flux-controlled model u = 0.5 µm min−1. The stochastic simulation uses the same
parameters except w = 1.5 µm min−1 and kend
off = 1 min−1 for the density-controlled model.
and the filament length is L(t). The mean-field density-dependent depolymerization model is:
(cid:32)
v − L
(cid:33)
pe =
L = u − wpe.
a
p(L − a, t)(1 − pe) − kend
off pe,
(4)
(5)
The first term on the right side of equation (4) represents the stepping of motors from the site
adjacent to the end to the terminal site on the filament, at rate (v − L)/a. The density at the
penultimate site on the filament is p(L− a, t), where a is the lattice spacing, assumed small. If the
motor dynamics at the end are faster than the typical time scale of density changes in the bulk
of the filament, we can treat the motor density away from the end quasi-statically assuming it is
unaffected by the end dynamics. Thus, we write p(L − a, t) ≈ p(L, t), where p(x, t) is the motor
occupancy for a region far from the filament end. Because the kinetics of motor removal may be
different at the end of the filament, we include crowding effects at the last site even though they
are neglected elsewhere along the microtubule. The second term on the right side of equation (4)
describes unbinding of the motor at the end. In equation (5), the filament lengthens at constant
speed u and shortens at a rate proportional to the motor density at the end, with a maximum
speed w.
Note that in this model the unbinding of the motor from end of the filament (controlled by the
off pe in equation (4)) is decoupled from the depolymerization rate (controlled by
term with rate kend
the term with rate wpe in equation (5)). This means that we allow processive depolymerization,
with a single motor able to remove an average of w/kend
off monomers from the filament.
The steady-state length in the density-controlled model Lden is reached when L = 0 and pe = 0,
6
which implies pe = u/w and
(cid:20)
(cid:18) ρmaxakend
1 − uakend
p0v(w − u)
off
(cid:19)
off
w − u
(cid:21)
(6)
(7)
Lden = −λ ln
≈ u
konc
The approximate solution in equation (7) applies when the second term inside the logarithm of
equation (6) is small. Note that there is no steady-state solution if either w < u (in this case the
off /(p0v(w −
motors can never remove dimers quickly enough to keep up with the growth) or uakend
off /(v(w−
u)) ≥ 1. Therefore the motor occupancy must be larger than the critical value p0c = uakend
u)) for a steady-state length to occur, implying a minimum bulk motor concentration of cc =
koff ρmaxakend
off ]). In practice given measured motor parameters, reasonable
values of the steady-state length require w quite similar to u; the requirement for such fine-tuning
of the depolymerization rate suggests that this length-regulation mechanism is not highly robust
(figure 2).
off /(kon[w(w/u− 1)− akend
We show the dependence of the steady-state length on the bulk motor concentration in figure
2. We use parameters similar to those found in experiments [11, 13]. The results of stochastic
simulation of density-controlled depolymerization agree qualitatively with the predictions of the
mean-field theory (figure 2, details of the simulations are described in section V). The best agree-
ment occurs when the stochastic simulation uses a slightly larger motor-induced depolymerization
speed w and a slightly lower motor unbinding rate from the filament end kend
off than the mean-field
theory.
In this case the mean filament lengths reached in the two models are similar, but the
stochastic simulation shows large fluctuations about the mean length. This is intuitively reason-
able since in this model depolymerization is controlled by the motor occupancy at the terminal
site of the filament, which undergoes significant fluctuations.
B. Flux-controlled depolymerization
Varga et al. found that their experimental data for depolymerization of stabilized microtubules
by Kip3p are consistent with filament depolymerization being determined by the flux of motors
to the end [13].
In this model, a motor would in principle remain bound to the filament end
forever, unless displaced from the tip by the arrival of another motor. When unbinding, each
motor is assumed to shorten the microtubule by a length δ (where δ could equal the lattice spacing
a if each motor removes exactly one tubulin dimers, but could differ from a depending on the
motor depolymerization processivity). Therefore the depolymerization speed is w = δJ(L), where
J(L) = p(L)ρmax(v− L) is the flux of motors to the end of the filament. Note that steric interactions
between motors that decrease the flux are neglected here. The length of the microtubule changes
in time according to
L = u − w = u − δρmaxp(L)
v − L
(8)
(cid:16)
(cid:17)
At steady state,
δρmaxvp(Lss), and the steady-state length in the flux-controlled model is
L = 0 and the motor occupancy is the steady-state value. Therefore u =
Lflux = −λ ln
≈ u
konc
1 −
(cid:19)
u
p0vδρmax
.
(cid:21)
(cid:20)
(cid:18) 1
δ
7
(9)
(10)
As above, the approximate solution applies when the second term inside the logarithm is small.
Note that there is no steady-state solution if u/(vδp0ρm) ≥ 1. This implies that the motor density
must be larger than the critical value p0c = u/(vδρmax) for a steady-state length to occur; this
corresponds to a critical bulk motor density cc = koff ρmaxu/(kon(vδρmax − u)). This requires
vδρmax > u; in practice, for parameters for the budding-yeast motor Kip3 vδρmax must be a few
times u to get steady-state lengths of a few microns.
We show the dependence of the steady-state length on the bulk motor concentration in figure 2.
The results of stochastic simulation of flux-controlled depolymerization agree quantitatively with
the predictions of the mean-field theory for identical parameters (figure 2, details of the simulations
are described in section V). Compared to the density-controlled depolymerization model, the flux-
controlled depolymerization model shows decreased fluctuations and a relatively narrow length
distribution. This may occur because in this model depolymerization is controlled by the motor
flux to end of the filament, which is a collective property of multiple motors.
The structures of the steady-state solutions are similar in the two models, having the approxi-
mate form (equations (7) and (10)) L ∼ u/konc times a factor with units of inverse length related
to how motors are removed from the end. These approximations to L make clear the strong depen-
dence of the steady-state length on the bulk motor concentration, implying that length regulation
by this mechanism requires tight regulation of the motor concentration c. In the density-controlled
model the motor unbinding rate from the end of the filament and the difference w − u between the
maximum speed of depolymerization and the filament growth rate are important in controlling the
length reached. In the flux-controlled model the steady-state length takes a simple form, depending
primarily on u, konc, and δ.
In both cases, there is a minimum motor occupancy required to reach a steady-state length,
as expected, because a minimum number of motors is required for depolymerization to balance
polymerization. The steady-state filament length is quite different from the dimensional length
scale λ which characterizes the motor density profile.
IV. LENGTH REGULATION BY ALTERING CATASTROPHE
In cells, microtubule filaments typically don't grow constantly as in the simple model above,
but instead undergo dynamic instability, characterized by long-lived growing and shrinking regimes
with transitions between these two states. Studies of Kip3p in cells [11] and in vitro [29] and of
other kinesin-8 motors [25, 35], as well as modeling work [37] suggest that in cells these proteins may
act to promote catastrophe (the transition from growing to shrinking) of dynamic microtubules.
Therefore, it is necessary to understand the consequences of length-dependent changes in filament
dynamics (rather than merely shortening) to predict the effects of these motors in cells.
Here we develop a theory of motors that promote filament catastrophe in a length-dependent
manner. The number probability density n(L) = nG(L) + nS(L) for filaments of length L is made
up of two populations, growing (G) and shrinking (S) filaments. The total number of filaments
is N =(cid:82) n(L) dL. In this model, we neglect pauses (neither growth nor shrinkage) exhibited by
dynamic microtubules in cells. The distributions evolve according to
∂nG
∂t
∂nS
∂t
= −u
= w
∂nG
∂L
∂nS
∂L
− fcnG + frnS
+ fcnG − frnS
8
(11)
(12)
The terms in the first equation represent filament growth with speed u, catastrophe with frequency
fc, and rescue with frequency fr. The terms in the second equation represent filament shortening
with speed w, catastrophe, and rescue. At steady state, u ∂nG
∂L . The solution to this
equation (consistent with the boundary condition that the number of filaments drops to zero as
L → ∞) is nS = (u/w)nG. Then the steady-state equation for the total number of filaments
n = nG + nS simplifies to
∂L = w ∂nS
(cid:18) fc
u
= −
∂n
∂L
− fr
w
(cid:19)
n
(13)
If the catastrophe and rescue rates are spatially constant, the microtubule length distribution is
exponential, n(L) = n0 exp(−(fc/u − fr/w)L), so the distribution is a bounded exponential if
fc/u > fr/w and has characteristic length uw/(fcw − fru).
In the case of length-dependent rates, we have the formal solution
(cid:90)
ln n = −
dL
(cid:19)
− fr
w
(cid:18) fc
u
(cid:90)
(cid:18)
− 1
u
(cid:19)
Here, we assume that only the catastrophe rate fc(L) varies with length (as observed for the
kinesin-8 motors Klp5/6 in fission yeast cells [25]), and other rates are all constant. Then
n = n0efrL/w exp
dL fc(L)
(15)
(14)
A. Density-controlled catastrophe
As above, we assume that the motors move faster than the filament growth, so the motors track
the end and the motor density is at steady state. In the density-controlled catastrophe model, we
assume that the catastrophe frequency increases linearly with the motor occupancy at the end of
the filament:
The motor occupancy at the end is determined by equation (4). At steady state, the occupancy
at the end of the growing filament is
f = fc + αpe.
(16)
pe =
b(1 − e−L/λ)
off + b(1 − e−L/λ)
kend
,
with b = (v − u)p0/a. Using the integral
(cid:90) L
0
dL(cid:48)
(1 − e−L(cid:48)/λ)
off + b(1 − e−L(cid:48)/λ)
kend
=
L
kend
off + b
− λkend
off
b(kend
off + b)
(17)
(cid:18)
ln
1 +
b
kend
off
(cid:19)
(1 − e−L/λ)
,
(18)
9
FIG. 3. Dynamics of length regulation by altering catastrophe. Top, density-controlled model. Bottom, flux-
controlled model. Left, example trace of filament length versus time in the stochastic simulation. Middle,
catastrophe frequency as a function of filament length, comparing mean-field theory (line) and stochastic
simulation (points). Right, length distribution of dynamic filaments for the model with and without motors,
comparing mean-field theory and stochastic simulation. The presence of the motors leads to a significant
decrease in the mean filament length. This figure uses the parameters v = 3 µm min−1, u = 1 µm min−1,
w = 7 µm min−1, minimum catastrophe frequency fc = 0.2 min−1, rescue frequency fr = 0.05 min−1,
a = 8 nm, and ρmax = 125 µm−1. For the density-controlled model the parameters are kon = 1 nM−1µm−1
min−1, koff = 0.25 min−1, kend
off = 1.5 min−1, bulk motor concentration c = 2 nM, and α = 0.35 min−1.
off = 1 min−1and α = 0.38
The simulation of the density-controlled model has the same parameters except kend
min−1. For the flux-controlled model the parameters are kon = 3 nM−1µm−1 min−1, koff = 0.25 min−1,
bulk motor concentration c = 4 nM, and α = 7 × 10−3. The simulation of the flux-controlled model has
the same parameters as the corresponding mean-field theory except kon = 1.5 nM−1µm−1 min−1 and
α = 2 × 10−3.
the length distribution is
n(L) = n0e−((fc+∆f )/u−fr/w)L
(cid:18)
1 +
(v − u)p0
akend
off
(cid:19)
(1 − e−L/λ)
αλkend
off
+(v−u)p0/a)
u(kend
off
,
(19)
off + (v − u)p0) is the maximum possible increase in the catastrophe
where ∆fden = α(v − u)p0/(akend
frequency in the density-controlled model. We see two effects due to motors: first, there is an
effective increase in the catastrophe rate of ∆fden. Second, there is an additional multiplicative
factor in the length distribution. This factor approaches one in the limit of short filaments (L (cid:28) λ)
and for typical experimental parameters varies slowly with L.
Note that in this model the unbinding of the motor from end of the filament is controlled by the
rate constant kend
off . This means that a motor can processively track the end of a depolymerizing
filament, and this processivity tends to increase motor concentration at the end of the filament and
therefore enhance the filament-shortening effects of motors.
B. Flux-controlled catastrophe
In the flux-controlled catastrophe model, we assume that the catastrophe frequency increases
linearly with the flux of motors to the end:
10
(cid:18) α(v − u)ρmaxp0λ
(cid:19)
(20)
The flux to the end of the microtubule is J = p(L)ρmax(v − u). Note that in this model α is
dimensionless and that steric interactions between motors that decrease the flux are neglected.
The length distribution is then
f = fc + αJ.
n = n0e−(fc/u+∆f /u−fr/w)L exp
(21)
where ∆fflux = α(v − u)ρmaxp0 is the maximum possible increase in the catastrophe frequency in
the flux-controlled model. Again, we see two effects from the length-dependent catastrophe: there
is an effective increase in the catastrophe rate of ∆fflux, and there is an additional multiplicative
factor in the length distribution which is an exponential of an exponential of the length. This
factor approaches 1 in the limit of short microtubules (L (cid:28) λ) and approaches the constant factor
exp
as L → ∞; for typical experimental parameters this factor is of order 1.
(cid:16) α(v−u)ρmaxp0λ
(cid:17)
u
(1 − e−L/λ)
,
u
We show simulations of filament length as a function of time, calculations of the variation
of catastrophe frequency with filament length and filament length distributions in figure 3. We
chose parameters from experiments on the increase in catastrophe frequency associated with the
kinesin-8 motor Klp5/6 in fission yeast[25], which found a catastrophe frequency fc = 0.2 min−1
in cells lacking kinesin-8 motors and a length-dependent increase in the catastrophe frequency up
to a maximum of 0.5 min−1 for filaments 8 µm long in cells containing motors. With the correct
choice of parameters, the length-dependent increase in catastrophe frequency due to motors is
qualitatively similar to that measured by Tischer et al. [25].
The results of stochastic simulation are shown for comparison with mean-field theory. For
density-controlled catastrophe, there is excellent agreement between stochastic simulation results
and mean-field theory if the parameters kend
off and α are slightly modified. The flux-controlled catas-
trophe model predictions of the dependence of catastrophe frequency on filament length show only
rough qualitative agreement with mean-field theory; the shapes of the curves are quite different.
Even this level of agreement requires modification of the parameters kon and α.
C. Mean filament length
The length-dependent catastrophe induced by motors changes the microtubule length distri-
bution in two ways: the effective catastrophe frequency increases, and the length distribution is
multiplied by an additional function. When the additional change in the functional form due to this
multiplication can be neglected, including only the effective increase in the catastrophe frequency
gives a simple result for the change in mean filament length. The approximate length distribution
is
(22)
We define ¯L0 = uw/(fcw − fru), the mean filament length in the absence of motors, and the mean
length including motor effects is ¯L = ¯L0 − ∆L. The mean length is
n(L) = n0e−[(fc+∆f )/u−fr/w]L.
(cid:18) fc − u
fc − u
w fr
w fr + ∆f
(cid:19)
¯L = ¯L0
(23)
11
FIG. 4. Mean filament length and changes in filament length as a function of motor parameters. Left,
variation as a function of bulk motor concentration. Center, variation as a function of motor speed. Right,
variation as a function of kend
off in the density-controlled model. This figure uses the same parameters as
figure 3 except where noted in varying the bulk motor concentration and motor speed: v = 3 µm min−1,
u = 1 µm min−1, w = 7 µm min−1, minimum catastrophe frequency fc = 0.2 min−1, rescue frequency
fr = 0.05 min−1, a = 8 nm, and ρmax = 125 µm−1. For the density-controlled model the parameters are
kon = 1 nM−1µm−1 min−1, koff = 0.25 min−1, kend
off = 1.5 min−1, bulk motor concentration c = 2 nM, and
α = 0.35 min−1. The simulation of the density-controlled model has the same parameters except kend
off = 1
min−1 and α = 0.38 min−1. For the flux-controlled mean-field model the parameters are the same as for the
density-controlled mean-field model except kon = 3 nM−1µm−1 min−1 and α = 7× 10−3. The simulation of
the flux-controlled model has the same parameters as the corresponding mean-field theory except kon = 1.5
nM−1µm−1 min−1 and α = 2 × 10−3.
The fractional change in the mean length is
∆L
¯L0
=
fc − u
∆f
w fr + ∆f
(24)
This expression is an approximation that typically overestimates the change in filament length
due to motors, but has the advantage that it has a simple analytical form that can be used
to understand which parameters control the mean length. The mean filament length is related
to the maximum increase in catastrophe frequency that can be achieved by the motors. In the
density-controlled model ∆fden = α(v − u)p0/(akend
off + (v − u)p0), and in the flux-controlled model
∆fflux = α(v − u)ρmaxp0.
The change in the mean filament length varies with the bulk motor concentration (through the
occupancy p0), the difference between the filament growth speed and the motor walking speed,
and the motor unbinding rate from the filament end in the density-controlled model (figure 4).
This suggests that in cells length regulation could be tuned by altering motor concentration, mo-
tor/filament velocity, or motor off rate at the end of the filament in the density-controlled model.
For typical experimental parameters, the fractional change in mean filament length varies from 0.1
to 0.9 with changes in these parameters. While changes in bulk motor concentration can alter the
mean filament length, relatively large changes in the concentration (over two orders of magnitude)
lead to only modest changes in mean filament length. We find that in comparison to stochastic
simulation the approximate expressions calculated above tend to overstate the changes in mean
12
filament length achievable through bulk motor density changes. Particularly for the flux-controlled
model, mean filament length is more sensitive to alterations in motor speed. Varying motor speed
over a factor of 4 allows approximately factor of 10 change in the mean filament length. Compar-
ison to stochastic simulations shows that mean filament length changes due to variation in motor
speed are accurately captured by the approximate model.
In the density-controlled model, the rate constant kend
off controls the residence time of a motor
at the end of a filament. In the limit that this rate becomes large compared to other rates in the
problem, the motors rapidly unbind upon reaching the filament end and the changes in filament
length due to motors become smaller.
The increase in catastrophe due to motors can shift the filament length distribution from the
unbounded growth regime to the bounded growth regime. If fr/w > fc/u, the filament dynamics
are in the unbounded growth regime, with no defined mean filament length.
If the increase in
catastrophe due to motors ∆f is large enough, the presence of motors can shift the distribution
back to the bounded growth regime characterized by an exponential length distribution and a
well-defined mean filament length. This requires that ∆f > fru/w − fc. This relationship implies
a minimum motor concentration to shift from unbounded to bounded growth.
V. STOCHASTIC SIMULATION
We developed a kinetic Monte Carlo simulation of the model as shown in figure 1. The model
considers a single filament (equivalent to representing a microtubule by a single protofilament)
made up of a varying number of monomers. Each motor occupies a single filament monomer.
Typical time scales of motor and filament processes are of order seconds to minutes, so we chose a
simulation time step of 0.01 s. At each time step a number of monomers equal to the total number
of monomers currently in the the filament is randomly sampled and one step of the polymerization
dynamics at the end of the filament is performed. Each site on the filament except the last site
has the same rules: a motor can bind to an empty site, if the next site forward is empty a motor
can step forward, and a motor can unbind to create an empty site.
The behavior of the last site (the end of the filament) varies depending on the model considered.
In the model of length regulation by depolymerization (figure 1A), the filament can grow by addition
of a monomer. Growth is independent of motor occupancy at the last site. The filament can
shrink by one monomer depending on the motor occupancy at the end of the filament.
In the
density-controlled depolymerization model, removal of the terminal monomer can occur if
the terminal site is occupied by a motor.
In this model the motor can processively track the
depolymerizing filament: if the penultimate site is empty, the motor at the last site steps backward
when the terminal monomer is removed. If the second-to-last site is occupied, the motor on the
terminal site is removed when the terminal monomer is removed. The motor at the terminal site
can also directly unbind from the filament without removal of a monomer. In the flux-controlled
depolymerization model, removal of the terminal monomer occurs when the last and penultimate
sites are both occupied by motors and the motor at the penultimate site attempts a forward step.
In this case the terminal motor and monomer are both removed.
In the model of length regulation by catastrophe (figure 1B), the filament stochastically switches
between growing and shrinking states. Speeds of growth and shrinkage as well as the rescue
frequency are independent of motor occupancy at the last site. The catastrophe frequency is
increased depending on the motor occupancy at the end of the filament. In the density-controlled
catastrophe model, the catastrophe frequency is increased by α if the terminal site is occupied by
a motor. In this model the motor can processively track the depolymerizing filament: if the filament
is shortening and the penultimate site is empty, the motor at the last site steps backward when the
13
terminal monomer is removed. If the second-to-last site is occupied, the motor on the terminal site
is removed when the terminal monomer is removed. The motor at the terminal site can also directly
unbind from the filament without removal of a monomer. In the flux-controlled catastrophe
model, the increase in the catastrophe frequency by α occurs when the last and penultimate sites
are both occupied by motors and the motor at the penultimate site attempts a forward step. In
this case the terminal motor is removed independent of whether or not a catastrophe occurs.
In all versions of the model, if the filament fully depolymerizes (0 sites remain) a new filament
is nucleated containing 1 site. For each parameter set we performed 5-10 simulations of 106 − 108
time steps.
VI. CONCLUSION
We have considered an example of biophysical length regulation by motors that walk along a
filament and promote filament shortening, inspired by experiments on the effects of kinesin-8 motor
proteins on microtubule dynamics. The motors bind to microtubules and move toward their plus
ends, and the presence of motors at the filament end alters microtubule polymerization dynamics.
The first mechanism we considered is a simplified model of length regulation, in which the
motors directly catalyze depolymerization of the filament from its plus end. When the action of
the motors is balanced by a constant filament polymerization rate, a steady-state filament length
can be reached. This mechanism neglects any fluctuating filament dynamics: only a single steady-
state length is reached. In the stochastic simulation, fluctuations due to stochastic motor/filament
dynamics lead to a spread about the mean length but the distribution of filament lengths remains
strongly peaked. There is a minimum bulk motor concentration on the filament to reach a steady-
state length, because if there are too few motors, motor-induced depolymerization can never be
fast enough to balance the intrinsic polymerization. In addition, inequalities involving the motor
motion constrain the parameter regime where steady-state solutions are possible. The steady-
state filament length differs from the length scale λ which characterizes the motor density profile.
The steady-state filament length depends sensitively on the bulk motor concentration, implying
that this mechanism of length regulation requires tight control of total motor number to operate
successfully.
Other recent theory work has addressed length regulation due to depolymerizing motors and
filament kinetics described by constant growth [39, 41] or treadmilling [40]. Govindan et al. con-
sidered a similar model for motor motion, but used an absorbing boundary condition for motors at
the filament plus end, an approximation that doesn't apply to filaments with biologically realistic
growth and shrinkage rates. Their work found an exponential filament length distribution for typi-
cal parameter values corresponding to Kip3 [39]. Melbinger et al. improved the model of Govindan
et al. by studying in detail how effects of motor crowding near the microtubule end control the
depolymerization dynamics. They discovered a parameter regime in which filament length is well
regulated, and how the length depends on motor kinetics [41]. Johann et al. considered the related
problem of how length regulation can be achieved by depolymerizing motors on filaments that
undergo treadmilling dynamics (addition of subunits at one end and removal at the other) [40].
In the model of length regulation by depolymerization we have discussed, constant growth is bal-
anced by length-dependent depolymerization. In the balance-point model of flagellar length regula-
tion in Chlamydomonas, a constant rate of flagellar disassembly is balanced by a length-dependent
rate of flagellar assembly, leading to a fixed flagellar length [10, 45]. While the underlying micro-
scopic mechanisms of flagellar length regulation differ from those discussed here, the conceptual
similarity is striking. Perhaps this basic idea of regulating length by making assembly or disas-
sembly length-dependent while the other process (disassembly or assembly) is length independent
14
could be a general paradigm for length regulation, at least of microtubule-based structures.
A more biologically relevant model for the length regulation of dynamic microtubules is length
regulation by altering catastrophe, in which the filament undergoes dynamic instability character-
ized by long-lived growing and shrinking states with transitions between growth and shrinkage. The
effect of the motors at the end is then not to directly depolymerize the filament but to increase
the catastrophe frequency. We calculate how the filament length distribution is altered by the
motor-dependent increase in catastrophe frequency, and derive a simple approximate expression
that relates the mean filament length to the maximum increase in catastrophe frequency that can
be achieved by the motors. The mean filament length varies modestly with bulk motor concentra-
tion but is sensitive to the difference between the filament growth speed and the motor walking
speed.
The increase in catastrophe frequency associated with the kinesin-8 motor Klp5/6 in fission
yeast cells was measured by Tischer et al. [25], who found a catastrophe frequency fc = 0.2 min−1
in cells lacking Klp5/6 and a length-dependent increase in the catastrophe frequency up to a
maximum of 0.5 min−1 for filaments 8 µm long in cells containing motors. With the correct choice
of parameters, our model displays a length-dependent increase in catastrophe frequency due to
motors which is qualitatively similar to that measured by Tischer et al. Using these parameters in
our model, changes in the mean length of a factor of 2 can be achieved by this mechanism.
ACKNOWLEDGMENTS
The authors thank Robert Blackwell, Matt Glaser, Loren Hough and Dick McIntosh for useful
discussions. This work was supported by NSF CAREER Award DMR-0847685, NSF MRSEC
Grant DMR-0820579, and NIH training grant T32 GM-065103.
REFERENCES
[1] S. J. Day and P. A. Lawrence. Measuring dimensions: the regulation of size and shape. Development,
127(14):2977 -- 2987, 2000.
[2] E. Hafen and H. Stocker. How are the sizes of cells, organs, and bodies controlled? PLoS Biology,
1(3):e86, 2003.
[3] Mike Cook and Mike Tyers. Size control goes global. Current Opinion in Biotechnology, 18(4):341 -- 350,
August 2007.
[4] G. Goshima, R. Wollman, N. Stuurman, J. M. Scholey, and R. D. Vale. Length control of the metaphase
spindle. Current Biology, 15(22):1979 -- 1988, 2005.
[5] Claire E. Walczak, Timothy J. Mitchison, and Arshad Desai. XKCM1: a xenopus kinesin-related protein
that regulates microtubule dynamics during mitotic spindle assembly. Cell, 84(1):37 -- 47, January 1996.
[6] F. Rivero, B. Koppel, B. Peracino, S. Bozzaro, F. Siegert, C. J. Weijer, M. Schleicher, R. Albrecht,
and A. A. Noegel. The role of the cortical cytoskeleton: F-actin crosslinking proteins protect against
osmotic stress, ensure cell size, cell shape and motility, and contribute to phagocytosis and development.
Journal of Cell Science, 109(11):2679 -- 2691, November 1996.
[7] C. Revenu, R. Athman, S. Robine, and D. Louvard. The co-workers of actin filaments:
from cell
structures to signals. Nature Reviews Molecular Cell Biology, 5(8):635 -- 646, 2004.
[8] Marileen Dogterom and Stanislas Leibler. Physical aspects of the growth and regulation of microtubule
structures. Physical Review Letters, 70(9):1347 -- 1350, March 1993.
[9] Anna Akhmanova and Casper C Hoogenraad. Microtubule plus-end-tracking proteins: mechanisms
and functions. Current Opinion in Cell Biology, 17(1):47 -- 54, February 2005.
15
[10] Wallace F. Marshall, Hongmin Qin, M´onica Rodrigo Brenni, and Joel L. Rosenbaum. Flagellar length
control system: Testing a simple model based on intraflagellar transport and turnover. Molecular
Biology of the Cell, 16(1):270 -- 278, January 2005.
[11] M. L. Gupta, P. Carvalho, D. M. Roof, and D. Pellman. Plus end-specific depolymerase activity of
Kip3, a kinesin-8 protein, explains its role in positioning the yeast mitotic spindle. Nature Cell Biology,
8(9):913 -- 923, 2006.
[12] V. Varga, J. Helenius, K. Tanaka, A. A. Hyman, T. U. Tanaka, and J. Howard. Yeast kinesin-8
depolymerizes microtubules in a length-dependent manner. Nature Cell Biology, 8(9):957 -- 962, 2006.
[13] Vladimir Varga, Cecile Leduc, Volker Bormuth, Stefan Diez, and Jonathon Howard. Kinesin-8 Motors
Act Cooperatively to Mediate Length-Dependent Microtubule Depolymerization. Cell, 138(6):1174,
2009.
[14] L. E. Hough, A. Schwabe, M. A. Glaser, J. R. McIntosh, and M. D. Betterton. Microtubule depolymer-
ization by the kinesin-8 motor kip3p: a mathematical model. Biophysical Journal, 96(8):3050 -- 3064,
2009.
[15] Robert R West, Terra Malmstrom, Cynthia L Troxell, and J R McIntosh. Two related kinesins, klp5+
and klp6+, foster microtubule disassembly and are required for meiosis in fission yeast. Molecular
Biology of the Cell, 12(12):3919 -- 32, 2001.
[16] Robert R West, Terra Malmstrom, and J Richard McIntosh. Kinesins klp5(+) and klp6(+) are required
for normal chromosome movement in mitosis. Journal of Cell Science, 115:931 -- 40, 2002.
[17] Miguel Angel Garcia, Nirada Koonrugsa, and Takashi Toda. Two kinesin-like Kin I family proteins in
fission yeast regulate the establishment of metaphase and the onset of anaphase A. Current biology,
12(8):610 -- 21, 2002.
[18] M. A. Garcia, N. Koonrugsa, and T. Toda. Spindle-kinetochore attachment requires the combined
action of Kin I-like Klp5/6 and Alp14/Dis1-MAPs in fission yeast. EMBO Journal, 21(22):6015 -- 6024,
2002.
[19] M. S. Savoian, M. K. Gatt, M. G. Riparbelli, G. Callaini, and D. M. Glover. Drosophila Klp67A is
required for proper chromosome congression and segregation during meiosis i. Journal of Cell Science,
117(16):3669 -- 3677, 2004.
[20] M. I. Mayr, S. Hummer, J. Bormann, T. Gruner, S. Adio, G. Woehlke, and T. U. Mayer. The human
kinesin Kif18A is a motile microtubule depolymerase essential for chromosome congression. Current
Biology, 17(6):488 -- 498, 2007.
[21] K. Jaqaman, E. M. King, A. C. Amaro, J. R. Winter, J. F. Dorn, H. L. Elliott, N. Mchedlishvili,
S. E. McClelland, I. M. Porter, M. Posch, et al. Kinetochore alignment within the metaphase plate
is regulated by centromere stiffness and microtubule depolymerases. The Journal of Cell Biology,
188(5):665 -- 679, 2010.
[22] R. Gandhi, S. Bonaccorsi, D. Wentworth, S. Doxsey, M. Gatti, and A. Pereira. The drosophila kinesin-
like protein KLP67A is essential for mitotic and male meiotic spindle assembly. Molecular Biology of
the Cell, 15(1):121 -- 131, 2004.
[23] M. K. Gatt, M. S. Savoian, M. G. Riparbelli, C. Massarelli, G. Callaini, and D. M. Glover. Klp67A
destabilises pre-anaphase microtubules but subsequently is required to stabilise the central spindle.
Journal of Cell Science, 118(12):2671 -- 2682, 2005.
[24] Amy Unsworth, Hirohisa Masuda, Susheela Dhut, and Takashi Toda. Fission yeast kinesin-8 Klp5 and
Klp6 are interdependent for mitotic nuclear retention and required for proper microtubule dynamics.
Molecular Biology of the Cell, 19(12):5104, 2008.
[25] Christian Tischer, Damian Brunner, and Marileen Dogterom. Force- and kinesin-8-dependent effects
in the spatial regulation of fission yeast microtubule dynamics. Molecular Systems Biology, 5, March
2009.
Current Biology, 20(4):374 -- 380, 2010.
[26] Y. Du, C. A. English, and R. Ohi. The kinesin-8 Kif18A dampens microtubule plus-end dynamics.
[27] Carsten Peters, Katju[scaron]a Brejc, Lisa Belmont, Andrew J. Bodey, Yan Lee, Ming Yu, Jun Guo,
Roman Sakowicz, James Hartman, and Carolyn A. Moores. Insight into the molecular mechanism of
the multitasking kinesin-8 motor. The EMBO Journal, 29(20):3437 -- 3447, September 2010.
[28] H. Wang, I. Brust-Mascher, D. Cheerambathur, and J.M. Scholey. Coupling between microtubule
sliding, plus-end growth and spindle length revealed by kinesin-8 depletion. Cytoskeleton, 67(11):715 --
728, 2010.
16
[29] Melissa K. Gardner, Marija Zanic, Christopher Gell, Volker Bormuth, and Jonathon Howard. Depoly-
merizing kinesins Kip3 and MCAK shape cellular microtubule architecture by differential control of
catastrophe. Cell, 147(5):1092 -- 1103, November 2011.
[30] Natsuko Masuda, Tetsuhiro Shimodaira, Shu-Jen Shiu, Noriko Tokai-Nishizumi, Tadashi Yamamoto,
and Miho Ohsugi. Microtubule stabilization triggers the plus-end accumulation of Kif18A/kinesin-8.
Cell Structure and Function, 36(2):261 -- 7, January 2011.
[31] Monika I. Mayr, Marko Storch, Jonathon Howard, and Thomas U. Mayer. A non-motor microtubule
binding site is essential for the high processivity and mitotic function of kinesin-8 Kif18A. PLoS One,
6(11):e27471, November 2011.
[32] Jason Stumpff, Yaqing Du, Chauca A. English, Zoltan Maliga, Michael Wagenbach, Charles L Asbury,
Linda Wordeman, and Ryoma Ohi. A tethering mechanism controls the processivity and kinetochore-
microtubule plus-end enrichment of the kinesin-8 Kif18A. Molecular cell, 43(5):764 -- 75, September
2011.
[33] X. Su, W. Qiu, M.L. Gupta Jr, J.B. Pereira-Leal, S.L. Reck-Peterson, and D. Pellman. Mecha-
nisms underlying the dual-mode regulation of microtubule dynamics by Kip3/kinesin-8. Molecular
cell, 43(5):751, 2011.
[34] L.N. Weaver, S.C. Ems-McClung, J.R. Stout, C. LeBlanc, S.L. Shaw, M.K. Gardner, and C.E. Walczak.
Kif18A uses a microtubule binding site in the tail for plus-end localization and spindle length regulation.
Current Biology, 2011.
[35] M. Erent, D. R. Drummond, and R. A. Cross. S. pombe kinesins-8 promote both nucleation and
catastrophe of microtubules. PloS One, 7(2):e30738, 2012.
[36] P. M. Grissom, T. Fiedler, E. L. Grishchuk, D. Nicastro, R. R. West, and J. R. McIntosh. Kinesin-8 from
fission yeast: A heterodimeric, plus-end -- directed motor that can couple microtubule depolymerization
to cargo movement. Molecular Biology of the Cell, 20(3):963 -- 972, 2009.
[37] Ludovic Brun, Beat Rupp, Jonathan J. Ward, and Fran¸cois N´ed´elec. A theory of microtubule catas-
trophes and their regulation. Proceedings of the National Academy of Sciences, 106(50):21173 -- 21178,
December 2009.
[38] Louis Reese, Anna Melbinger, and Erwin Frey. Crowding of molecular motors determines microtubule
depolymerization. Biophysical Journal, 101(9):2190 -- 2200, November 2011.
[39] B. S. Govindan, M. Gopalakrishnan, and D. Chowdhury. Length control of microtubules by depoly-
merizing motor proteins. Europhysics Letters, 83(4):40006, August 2008.
[40] Denis Johann, Christoph Erlenkamper, and Karsten Kruse. Length regulation of active biopolymers
by molecular motors. Physical Review Letters, 108(25):258103, June 2012.
[41] Anna Melbinger, Louis Reese, and Erwin Frey. Microtubule length regulation by molecular motors.
Physical Review Letters, 108(25):258104, June 2012.
[42] Christian Tischer, Pieter Rein ten Wolde, and Marileen Dogterom. Providing positional information
with active transport on dynamic microtubules. Biophysical Journal, 99(3):726 -- 735, August 2010.
[43] A. Parmeggiani, T. Franosch, and E. Frey. Totally asymmetric simple exclusion process with Langmuir
kinetics. Physical Review E, 70(4), October 2004.
[44] Sarah A. Nowak, Pak-Wing Fok, and Tom Chou. Dynamic boundaries in asymmetric exclusion pro-
cesses. Physical Review E, 76(3):031135 -- 11, 2007.
[45] Benjamin D. Engel, William B. Ludington, and Wallace F. Marshall. Intraflagellar transport particle
size scales inversely with flagellar length: revisiting the balance-point length control model. The Journal
of Cell Biology, 187(1):81 -- 89, October 2009.
|
1104.0876 | 1 | 1104 | 2011-04-05T16:06:29 | Direct measurement of the correlated dynamics of the protein-backbone and proximal waters of hydration in mechanically strained elastin | [
"physics.bio-ph",
"q-bio.SC"
] | We report on the direct measurement of the correlation times of the protein backbone carbons and proximal waters of hydration in mechanically strained elastin by nuclear magnetic resonance methods. The experimental data indicate a decrease in the correlation times of the carbonyl carbons as the strain on the biopolymer is increased. These observations are in good agreement with short 4ns molecular dynamics simulations of (VPGVG)3, a well studied mimetic peptide of elastin. The experimental results also indicate a reduction in the correlation time of proximal waters of hydration with increasing strain applied to the elastomer. A simple model is suggested that correlates the increase in the motion of proximal waters of hydration to the increase in frequency of libration of the protein backbone that develops with increasing strain. Together, the reduction in the protein entropy accompanied with the increase in entropy of the proximal waters of hydration with increasing strain, support the notion that the source of elasticity is driven by an entropic mechanism arising from the change in entropy of the protein backbone. | physics.bio-ph | physics | Direct measurement of the correlated dynamics of the protein-backbone and proximal
waters of hydration in mechanically strained elastin
Cheng Sun, Odingo Mitchell, Jiaxin Huang and Gregory S. Boutis∗
(Dated: June 6, 2021)
We report on the direct measurement of the correlation times of the protein backbone carbons
and proximal waters of hydration in mechanically strained elastin by nuclear magnetic resonance
methods. The experimental data indicate a decrease in the correlation times of the carbonyl carbons
as the strain on the biopolymer is increased. These observations are in good agreement with short
4ns molecular dynamics simulations of (VPGVG)3, a well studied mimetic peptide of elastin. The
experimental results also indicate a reduction in the correlation time of proximal waters of hydration
with increasing strain applied to the elastomer. A simple model is suggested that correlates the
increase in the motion of proximal waters of hydration to the increase in frequency of libration of
the protein backbone that develops with increasing strain. Together, the reduction in the protein
entropy accompanied with the increase in entropy of the proximal waters of hydration with increasing
strain, support the notion that the source of elasticity is driven by an entropic mechanism arising
from the change in entropy of the protein backbone.
PACS numbers: 87.64.kj, 87.80.Lg
Elastin, the principal protein component of the elas-
tic fiber, is a remarkable biopolymer that gives rise to
the resilience of many vertebrate tissues
[1 -- 3]. The
structure-function relationship of this system has been
of interest since Partridge's original work in isolating
elastin from other tissue constituents and a model in-
volving cross-links of hydrophobic and hydrophilic do-
mains [4]. Central to the elasticity of elastin is the de-
gree of protein solvation as well as the polarity of the
solvent as made evident by experimental studies of the
Young's modulus -- elastin only exhibits elasticity when
hydrated [5]. A 'liquid drop elastomer' model was sug-
gested wherein both the surface interaction between hy-
drophobic and hydrophilic phases and the change in con-
figurational entropy contribute to elasticity [6]. The no-
tion that the elasticity of elastin is entropic in origin was
established early on and a model was proposed that in-
corporated a network of random chains
[7]. However,
experimental work gave evidence of ordered structures in
elastin, specifically, the presence of both α-helices and
β-turns [1, 8]. The short model peptide (VPGVG)n, an
abundant motif of elastin, has been extensively studied
as it exhibits the inverse temperature transition that is
characteristic of the hydrated protein [1, 9]. In studies
of the (VPGVG)n peptide, a model was proposed that
incorporates a hydrogen bond formed between Val1 and
Val4 residues, resulting in suspended segments between
β-turns [10]. In this model, termed the Librational En-
tropy Mechanism (LEM), the suspended segments are
free to undergo large-amplitude low-frequency torsional
oscillations; the amplitude might be damped during ex-
tension, giving rise to a decrease in entropy of the seg-
ment with increasing strain, providing the driving mecha-
nism for elasticity [10]. In addition, type II β-turns have
∗ Email:[email protected], Phone: (718)951-5000x2873,
Fax: (718)951-4407, Physics Department, Brooklyn College of
The City University of New York, Brooklyn NY 11210
also been proposed for the (GXGGX) sequence, a sec-
ond repeating motif of elastin, wherein hydrogen bonds
are formed either between Gly1 and Gly4 residues or be-
tween the second and fifth X residues; the interchange
between the two conformations results in a dynamical β-
turn sliding giving rise to elasticity [2]. In both model
sequences, according to the LEM model, the elastic force
is borne by the protein backbone by virtue of a decrease
in entropy upon extension.
In all elastin models, wa-
ters of hydration are tacitly assumed and yet the contri-
bution of proximal waters of hydration to the entropic
mechanism of elasticity has eluded experimentalists. Re-
cent short time Molecular Dynamics (MD) simulations
of (VPGVG)18 have suggested that the orientational en-
tropy of waters hydrating hydrophobic groups decreases,
whereas the elastin backbone is more dynamic and has
a greater entropy upon extension than in a relaxed state
in disagreement with the LEM model [9]. In this work
we report on the first direct measurement of the motional
correlation times of proximal waters of hydration and the
protein backbone of mechanically strained elastin.
Nuclear Magnetic Resonance (NMR) experiments were
performed on purified bovine nuchal
ligament elastin
purchased from Elastin Products Company, LLC. The
elastin was purified by the neutral extraction method of
Partridge [4], and was suspended in a mixture of D2O
and H2O with a volume ratio of 50:50 with 0.0003g/mL
of sodium azide added as a biocide. The elastin samples
were mechanically strained in 1.5cm long, 5mm diameter
NMR tubes with both ends held by polytetrafluoroethy-
lene tape and sealed using ethylene-vinyl acetate. Four
samples were prepared; the strain (∆L/L) on the samples
was 0, 17, 35 and 43 percent, and are denoted I, II, III,
and IV respectively in the remainder of this work. Addi-
tionally, after the experiments were performed on sample
IV, the sample was released and measured after 72 hours
and is denoted sample V. The loss of water in any of the
samples was less than 1 percent over the entire course of
1
1
0
2
r
p
A
5
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
6
7
8
0
.
4
0
1
1
:
v
i
X
r
a
experiments. All the experiments were carried out on a
200MHz Tecmag Apollo NMR spectrometer using a Doty
solids NMR probe. The 90o pulse time was 9µs in the
13C measurements and 19µs in the 2H measurements. In
the 13C measurements, a radio frequency field strength
of ωe=27.7kHz was applied with 1H SPINAL-64 decou-
pling during signal acquisition [11]. All the experiments
were conducted at approximately 25oC.
We employed the 2D T1-T2 NMR technique [12], and
measured the 2H T1 and T2 relaxation times of water
in hydrated elastin. By applying a 2D Inverse Laplace
Transform (ILT) of the experimental data, the T1 and T2
relaxation times are correlated and are manifested into
a peak in the 2D map. The 2D ILT map of sample I is
shown in Fig. 1 as an example. Four peaks are observed
in the figure, corresponding to four reservoirs of water
with distinguishable dynamical properties; the 2D ILT
maps for all samples studied showed four components.
For a spin I=1 nucleus, such as 2H, the relaxation is
governed by the quadrupolar interaction and the T1 and
T2 are given by
1
T1
=
3
40
(1 +
η2
3
)Cq{J(ωD) + 4J(2ωD)}
(1)
1
T2
=
1
80
(1 +
η2
3
)Cq{9J(0) + 15J(ωD) + 6J(2ωD)} (2)
τc
where ωD is the Larmor frequency for 2H, Cq = ( eQ ∂2V
∂z2 )2
and η is the asymmetry parameter of the potential V and
assumed to 0 in our case. The spectral density is given
by J(ω) =
1+(ωτc)2 with τc defined as the correlation
time of the fluctuating quadrupolar field of the 2H nu-
cleus [13]. Using Eq. 1 and 2 and our measured values of
T1 and T2, the τc was determined for the four components
shown. Component α1 is free water, by virtue of the fact
that T1 ≈ T2, α2 is water that resides between the elastin
fibers and β is a mobile component that resides within
the elastin fiber as verified by a T2-T2 exchange experi-
ment [14]. The water corresponding to the peak with the
shortest T1 and T2 relaxation times, labeled γ in Fig. 1,
has been determined to be in closest proximity to elastin,
also by the T2-T2 exchange experiment and by consider-
ing the correlation times of the slowest two components
in the 2D ILT map (τ β
c =111.7ns). In this
work, we therefore study the changes in the correlation
time of component γ as a function of mechanical strain.
For samples I, II, III and IV, the correlation times are
plotted in Fig. 2a as a function of ∆L/L. It is evident
from the figure that the correlation time of proximal wa-
ters of hydration decreases with increasing strain on the
biopolymer. Not shown in the figure is the correlation
time of the water from sample V, which was measured to
be the same, within our experimental uncertainty, as that
measured in sample I. This observation indicates that the
dynamics of the water in close proximity to the protein
c =48.8ns and τ γ
2
FIG. 1. 2H 2D ILT map of the T1-T2 NMR relaxation times
of water in sample I, defined in the text. Four components
are discernable and are labeled α1, α2, β and γ. The dashed
lines are used to guide the eye for the region of the 2D map
where T1 is approximately equal to T2. The signal intensity,
indicated by the colorbar, is shown on a logarithmic scale.
FIG. 2. Correlation times measured in samples I, II, III and
IV as defined in the text. (a) Correlation time of the fluctuat-
ing quadrupolar field of D2O in close proximity to elastin, as
determined from Eq. 1 and 2 and the measured 2H T1 and T2.
(b) Correlation times characterizing the fluctuating dipolar
field experienced by the carbonyl (open circles) and aliphatic
carbons (closed circles) of elastin, as determined from Eq. 3
and 4 and the measured 13C T1 and T1ρ. The dashed lines
are used to guide the eye.
returned to what was observed in an unstrained sample
when the applied strain was removed.
Natural abundance 13C Direct Polarization (DP) and
1H → 13C Cross Polarization (CP) NMR experiments
were performed [15], and exemplary spectra are shown
in Fig. 3 from sample I. A phase cycling scheme was
implemented in the CP experiment, such that the result-
ing signal observed on the 13C spectra results only from
cross-polarized magnetization. Two features are evident
in the spectra. First, it is clear from the DP spectrum,
shown in Fig. 3a, that the signal for the carbonyl car-
bons (centered at 173ppm) is clearly distinguishable from
that of the aliphatic carbons (16ppm-60ppm) even with-
out any magic angle spinning [16]; this allows us to dis-
tinguish these two signals and measure the respective re-
laxation times unambiguously. Second, the overall signal
in the CP spectrum (Fig. 3b) is smaller than that of the
DP spectrum (Fig. 3a), indicating a high degree of mo-
γβα2α10510152025303540456080100120∆L / L [%]τc [ns] 051015202530354045400600800∆L / L [%]τc [ns] ab3
that the correlation times for both carbonyl and the over-
all aliphatic carbons decrease with increasing strain. The
physics of this result dictates an increase in the fluctua-
tion of the dipolar field surrounding both carbonyl and
aliphatic regions with increasing strain. The measured
correlation times of the carbonyl and overall aliphatic
carbons of sample V, not shown, were within the experi-
mental uncertainty of that observed in sample I. To help
visualize possible dynamical properties of strained elastin
measured in our experimental work, we have studied the
elastin mimetic sequence (VPGVG)3 by MD simulation.
The MD simulations were performed at 25oC using the
OPLS-AA/L force field model [18] and a Berendsen ther-
mostat [19] in GROMACS [20]. The peptide was termi-
nated by (-NH-CH3) and (-CO -- CH3) groups at its C- and
N-terminus, respectively. The model system was placed
in a cubic box solvated with water using the SPC216 wa-
ter model
[21] and was run for 4ns. Two scenarios were
studied: (a) without an applied pull force and (b) with
the Gly14 atoms being fixed and a pull force being ap-
plied on the atoms of Val4 with a constant force equal to
50000 kJ/mol nm. The resulting increase in the radius
of gyration, as determined by the position of the Cα, is
highlighted in Table 1 as well as the increase in the num-
ber of water molecules in 0.6nm of the peptide resulting
from the applied strain. Further details of the mechan-
ical properties of this elastin mimetic peptide, and oth-
ers, will appear in a forthcoming manuscript. For each
scenario the eigenvalues, λii, of the mass weighted co-
variance matrix of the MD trajectory were computed. A
quasi-harmonic approximation was applied whereby the
frequencies, ωi of uncorrelated simple harmonic oscilla-
tors are given by ωi =
[22]. The entropy of a
harmonic oscillator with frequency ωi at a temperature
T is given by
(cid:113) kB T
λii
Si = kB[
ωi/kBT
eωi/kB T − 1
− ln(1 − e−ωi/kB T )]
(5)
The histogram of the frequencies derived from the simu-
lations are plotted in Fig. 4 for both scenarios studied.
The resulting total entropy of the peptide was determined
by summing the entropy contribution of each frequency
and is tabulated in Table 1. Fig. 4 shows a shift in pop-
ulation to higher frequency when the peptide is strained.
According to Eq. 5 high frequencies contribute less to en-
tropy, thus Fig. 4 implies a decrease in the total peptide
entropy when strained; this is observed in our results tab-
ulated in Table 1. In the LEM model [10], the peptide
is treated as a series of quasi-harmonic oscillators un-
dergoing low-frequency, large torsional oscillations when
relaxed; the amplitude of the motion is reduced upon ex-
tension. Our experimental data, highlighted in Fig. 2b,
indicate that there is a decrease in the correlation times
of both carbonyl and aliphatic carbons, implying that the
motion of the fluctuating dipolar field of both regions in-
creases with increasing strain. We attribute the decrease
in the measured correlation times to an increase in fre-
quency of the protein backbone motion as observed in our
FIG. 3. Natural abundance 13C NMR spectra of sample
I defined in the text. (a) Direct Polarization (DP) and (b)
1H → 13C Cross Polarization (CP). The contact time for 1H
→ 13C CP was 1.4ms. The data shown were accumulated
with 14512 scans, with a recycle delay of 5s. All spectra are
referenced to adamantane.
bility for this system (a factor of 3.977 in gain is expected
for cross polarization for the case of a rigid solid [15]).
Similar spectral resolution was achieved on all samples
studied.
Using the DP technique, we obtained the T1 and T1ρ
times for the carbonyl and aliphatic carbons of mechani-
cally strained elastin. Given the low spectral resolution of
the static sample, one cannot distinguish different signals
in the aliphatic region and the dynamics of individual
constituents may vary. However, the 13C T1 and 1H T1ρ
values measured on hydrated elastin using magic angle
spinning methods at a similar temperature showed sim-
ilar values for the different components of the aliphatic
region [16]. It is therefore reasonable to study the region
by employing an effective T1 and T1ρ relaxation time. In
hydrated elastin, it has been demonstrated that the car-
bonyl chemical shift anisotropy is averaged to a great
extent due to the high mobility of the system [16]. The
13C relaxation times are thus mediated largely by carbon-
proton dipolar interactions and the 13C T1 and T1ρ are
given by
1
T1
=
1
10
CD{J(ωH − ωC) + 3J(ωC) + 6J(ωH + ωC)} (3)
1
T1ρ
CD{4J(ωe) + J(ωH − ωC) + 3J(ωC)
=
1
20
2γC γH
r3
C−H
+6J(ωH ) + 6J(ωH + ωC)} (4)
where ωH and ωC are the Larmor frequencies for 1H and
13C respectively, and CD = (
)2 [17]. In the spec-
tral density J(ω) in Eq. 3 and 4, τc is the correlation
time of the fluctuating dipolar field experienced by the
13C spins, which is reflective of the overall motions of the
13C and nearby 1H nuclei. Using Eq. 3 and 4, and the
measured relaxation times the τc is determined for the
carbonyl carbons and aliphatic carbons in all the sam-
ples. The results are presented in Fig. 2b as a function
of the strain on the sample. It is clear from the figure
−50050100150200−202468ppmIntensity [a.u.]−50050100150200−202468ppmIntensity [a.u.]ab4
sulting from the strain, as quantified by the Root Mean
Square Fluctuation (RMSF) of the Cα shown in Table 1,
is also consistent with the expected change according to
the LEM model.
Referring to Fig. 2a, the measured decrease in corre-
lation time indicates an increase in motion and hence an
increase in entropy of the proximal waters of hydration
with increasing strain. This observation appears to be
correlated with the observation that the lifetime of hydro-
gen bonds formed between the model peptide (VPGVG)3
and water decreases upon extension, as noted in Table
1. Given that the deformation is applied to the protein
backbone, it is clear that the changes to the dynamics of
the proximal waters of hydration are driven by the in-
crease in the frequency of libration of the backbone. The
increase in the entropy of the proximal waters of hydra-
tion and reduction in entropy of the backbone upon me-
chanical strain therefore suggests that the driving mech-
anism of elasticity is a change in the backbone entropy.
While the experimental data suggest that the proximal
waters of hydration in elastin increase mobility upon in-
creasing strain, the presence of water is a required, albeit
complex element for the elasticity of elastin. Ultimately
the solvent gives rise to backbone mobility whose increase
in librational frequency with increasing strain reduces the
backbone entropy and drives the system back to a relaxed
state.
The authors thank Yi-Qiao Song for use of the 2D ILT
algorithm and Raymond Tung, Alexej Jerschow, Rana-
jeet Ghose and Nicolas Giovambattista for useful discus-
sions. G. S. Boutis acknowledges support from NIH grant
number 7SC1GM086268-03. The content is solely the re-
sponsibility of the authors and does not necessarily repre-
sent the official views of the National Institute of General
Medical Sciences or the National Institutes of Health.
FIG. 4. Histogram of frequencies derived from the quasi-
harmonic approach in the MD simulations of the elastin
mimetic peptide (VPGVG)3 under relaxed (blue line) and
strained (red line) states.
Relaxed
Strained
Peptide entropy [kJ/mol·K]
1.63
1.46
1132±248 5451±177
No. of H2O within 0.6nm of peptide
0.19±0.05 0.14±0.06
Average RMSF of Cα [nm]
0.78±0.07 0.83±0.03
Radius of gyration of Cα [nm]
Lifetime of petpide-water H-bond [ps] 0.87±0.05 0.53±0.02
TABLE I. Molecular Dynamics (MD) simulation results for
an elastin mimetic peptide (VPGVG)3 under relaxed and
strained states, as described in the text. Peptide-water H-
bonds were counted when the donor-acceptor cutoff distance
of 0.35nm and the hydrogen-donor-acceptor angle of 30o was
satisfied.
4ns MD simulations. The observed increase in backbone
motion is also consistent with resulting MD simulations
of strained (VPGVG)18
[9]. In addition, the observed
reduction of the amplitude of the backbone motion re-
[1] D.W. Urry, J. Protein Chem. 7, 1 (1988).
[2] L. Debelle and A.M. Tamburro, Int. J. Biochem. Cell
Biol. 31, 261 (1999).
[3] L.D. Muiznieks, A.S. Weiss and F.W. Keeley, Biochem.
Cell Biol. 88, 239 (2010).
261 (2002).
[13] A. Abragam, P rinciples of N uclear M agnentism (Ox-
ford University Press, 1961).
[14] C. Sun and G.S. Boutis, New J. Phys. 13 025026 (2011).
Introduction to Solid − State N M R
[15] M.J. Duer,
[4] S.M. Partridge, H.F. Davies and G.S. Adair, Biochem. J.
Spectroscopy (Blackwell Publishing Ltd., 2004).
61, 11 (1955).
[5] M.A. Lillie and J.M. Gosline, J. Biorheol. 30, 229 (1993).
[6] T. Weis-Fogh and S.O. Anderson, Nature 227, 718
[16] A Perry, M.P. Stypa, B.K. Tenn and K.K. Kumashiro,
Biophys. J. 82, 1086 (2002).
[17] J.W. Peng, V. Thanabal and G. Wagner, J. Magn. Reson.
(1970).
[7] C.A.J. Hoeve and P.J. Flory, Biopolymers 13, 677 (1974).
[8] M. Mammi, L. Gotte and G. Pezzin, Nature 225, 371
(1968).
94, 82 (1991).
[18] W.L. Jorgensen and J. Tirado-Rives, J. Am. Chem. Soc.
110, 1657 (1988).
[19] H.J.C. Berendsen, J.P.M. Postma, A. DiNola and J.R.
[9] B. Li, D.O.V. Alonso, B.J. Bennion and V. Daggett.
Haak, J. Chem. Phys. 81, 3684 (1984).
2001. J. Am. Chem. Soc. 123, 11991 (2001).
[20] B. Hess, C. Kutnzer, D.V.D. Spoel and E. Lindahl, J.
[10] D.W. Urry and T.M. Parker, J. Muscle Res. Cell Motil.
Chem. Theory Comput. 4, 435 (2008).
23, 543 (2002).
[21] O. Teleman, B. Jonsson and S. Engstrom, Mol. Phys. 60,
[11] B. M. Fung, A.K. Khitrin and K. Ermolaev, J. Magn.
193 (1987).
Reson. 142, 97 (2000).
[22] I. Andricioaei and M. Karplus, J. Chem. Phys. 115, 6289
[12] Y.-Q. Song, L. Venkataramanan, M.D. Hurlimann, M.
Flaum, P. Frulla and C. Straley, J. Magn. Reson. 154,
(2001).
5101520253035404550010203040506070Frequency [MHz]Population |
1204.1418 | 1 | 1204 | 2012-04-06T06:40:26 | Multiple barriers in forced rupture of protein complexes | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.BM"
] | Curvatures in the most probable rupture force ($f^*$) versus log-loading rate ($\log{r_f}$) observed in dynamic force spectroscopy (DFS) on biomolecular complexes are interpreted using a one-dimensional free energy profile with multiple barriers or a single barrier with force-dependent transition state. Here, we provide a criterion to select one scenario over another. If the rupture dynamics occurs by crossing a single barrier in a physical free energy profile describing unbinding, the exponent $\nu$, from $(1- f^*/f_c)^{1/\nu}\sim(\log r_f)$ with $f_c$ being a critical force in the absence of force, is restricted to $0.5 \leq \nu \leq 1$. For biotin-ligand complexes and leukocyte-associated antigen-1 bound to intercellular adhesion molecules, which display large curvature in the DFS data, fits to experimental data yield $\nu<0.5$, suggesting that ligand unbinding is associated with multiple-barrier crossing. | physics.bio-ph | physics |
Multiple barriers in forced rupture of protein complexes
Changbong Hyeon1 and D. Thirumalai2
1Korea Institute for Advanced Study, Seoul 130-722, Republic of Korea
2Biophysics Program, Institute for Physical Science and Technology,
University of Maryland, College Park, MD 20742, USA
Curvatures in the most probable rupture force (f∗) versus log-loading rate (log rf ) observed in dy-
namic force spectroscopy (DFS) on biomolecular complexes are interpreted using a one-dimensional
free energy profile with multiple barriers or a single barrier with force-dependent transition state.
Here, we provide a criterion to select one scenario over another. If the rupture dynamics occurs
by crossing a single barrier in a physical free energy profile describing unbinding, the exponent ν,
from (1 − f∗/fc)1/ν ∼ (log rf ) with fc being a critical force in the absence of force, is restricted to
0.5 ≤ ν ≤ 1. For biotin-ligand complexes and leukocyte-associated antigen-1 bound to intercellular
adhesion molecules, which display large curvature in the DFS data, fits to experimental data yield
ν < 0.5, suggesting that ligand unbinding is associated with multiple-barrier crossing.
PACS numbers: 87.10.-e,87.15.Cc,87.80.Nj,87.64.Dz
INTRODUCTION
Single molecule pulling experiments have generated a
wealth of data, which can be used to probe aspects of
folding that were not previously possible [1 -- 3].
In ad-
dition, DFS has been used to decipher the energy land-
scape of molecular complexes by measuring the rupture
force (f ) by linearly increasing load at a rate rf (=
df /dt). Because of the stochastic nature of the unbind-
ing events, f varies from one complex (or realization)
to another, giving rise to an rf -dependent rupture force
distribution (P (f )). For a molecular complex obeying
Bell's formula, k(f ) = kof f exp (f x‡/kBT ), Evans and
Ritchie showed that the most probable force is f∗ =
(kBT /x‡) log(rf x‡/kof f kBT ) [4], where x‡(= xts − xb) is
the location of the transition state (xts) from the bound
state (xb) projected along the pulling coordinate and kof f
is the unbinding rate in the absence of force. However,
Bell's formula is applicable only if the molecular com-
plexes are mechanically brittle or if the applied tension
is sufficiently small that x‡ does not shift upon applica-
tion of force [5]. More generally, f∗ follows a (log rf )ν
dependence where ν depends on the details of the as-
sumed one dimensional (1D) model potential [6 -- 13]. The
basic assumption in all these works is that a single free
energy barrier along the pulling coordinate is sufficient
to describe force-driven rupture of the bound complex.
Sometime ago Merkel et al. used DFS to probe load de-
pendent strength of biotin bound to ligands, streptavidin
and avidin [14], showing that over six orders of variation
in rf (from about 10−2 to rf in excess of 104 pN/s) the
plot of f∗ versus log rf ([f∗, log rf ] plot) varies nonlin-
early for both ligands. We note parenthetically that it
is also common to observe curvature in unfolding rates
of proteins when the rf is varied [15]. By careful data
analysis combined with molecular dynamics simulations
they proposed an energy landscape for the complex, with
multiple energy barriers [14]. A similar picture emerges
in the rupture of intercellular adhesion molecules (ICAM-
1 and ICAM-2) bound to leukocyte function-associated
antigen-1 (LFA-1) upon application of force [16].
In principle, however, nonlinearity in [f∗, log rf ] plot
could also arise from load dependent variation in x‡
[17] in a 1D energy landscape with a single barrier [5 --
10, 12, 13, 17]. A theoretical model describing force-
induced escape from a bound state with a single barrier
in a cubic potential (ν = 2/3) has been used to ratio-
nalize the biotin-ligand data by identifying various linear
regimes demarcated by rf [9]. However, in the absence of
easily discernible changes in the slopes in [f∗, log rf ] plot
it is difficult to justify such an analysis. Here, we show
by analyzing experimental data that the observed non-
linearity in the DFS data of several protein complexes can
be better accounted for with an energy landscape con-
taining multiple sequential barriers, as originally demon-
strated [14, 16].
To illustrate how steep curvatures in DFS data can
arise naturally from a 1D free energy profile we cal-
culated P (f ) and [f∗, log rf ] of forced-escape kinetics
of a quasiparticle from a potential with two barriers,
U (x) = Ax(x − 1)[(x − 2)(x − 3)(x − 4)(x − 5) + 1] with
A > 0 (Fig.1). The distributions P (f ) are typical of
what is observed in experiments (Fig.1b). For all val-
ues of A, [f∗, log rf ] plots are curved although one could
discern a modest change in slope (Fig.1c). The load-
ing rate dependent x‡(rf ), calculated from the slope of
the data kBT /x‡(rf ) at each rf in Fig.1c, changes from
∼ 3 nm to < 1 nm. The precipitous change in x‡ at
rf ≈ (e−3 − e0) pN/s reflects the transition of the confin-
ing barrier from outer to inner barrier with an increasing
force (see Fig.1a). In contrast, gradual change of x‡ in
the range 0 < log rf < 10 is most likely due to the move-
2
FIG. 2. The n-dependent shape of G(x) (Eq.1). The poten-
tial with increasing n is associated with more brittle molecular
complexes. The yellow circle (x = xc) denotes an inflection
point and a cusp in each even and odd n potential, respec-
tively.
tial with a single barrier. Consider a Kramers' problem
of barrier crossing in a free energy profile G(x) in which
a single barrier separates the bound and unbound states
of a quasi-particle as in ligand bound in a pocket of a
receptor:
G(x) = G(xc) + fc(x− xc) +
(−1)n+1M
(n + 1)!
(x− xc)n+1 (1)
with M > 0. In G(x), a 1D free energy profile with a sin-
gle barrier, the shape of barrier and energy well is approx-
imated using n-th order polynomial with n = 1, 2, 3,··· .
For odd n, we assume that G(x) = −∞ for x > xc, so
that the transition state of G(x) is cusped. In the ab-
sence of tension, the barrier height, G‡, and the location
of transition state, x‡, are G‡ = χ n
n+1 fc(n!fc/M )1/n and
x‡ = χ(n!fc/M )1/n, respectively, where χ = 1 (for odd
n G‡/x‡. The form
n), 2 (for even n). Thus fc = n+1
of G(x), an extension of the microscopic theories using
harmonic-cusp or linear-cubic potential, accounts for the
degree of plasticity (or ductility) or brittleness of the
energy landscape [4] by changing n (Fig.2) [5]. Under
tension, Gef f (x) = G(x) − f · x; fc should be replaced
with fc(1 − f /fc) = fcε. Therefore, G‡(f ) = G‡ε1+1/n
and x‡(f ) = x‡ε1/n. Although Eq.1 looks similar to
the one Lin et al. used to discuss rupture dynamics
for ε (cid:28) 1 where the barrier height is almost negligi-
ble [12], we did not impose any specific force condition
on G(x). Instead of attributing the movement of tran-
sition state to a large external tension [7 -- 10, 12, 13],
we mapped the non-linearity in DFS data onto G(x)
that has the n-dependent shape of transition barrier and
bound state. In G(x), increasing brittleness makes x‡(f )
insensitive to applied tension, which is dictated by n;
x‡(f )/x‡ = ε1/n → 1. For a generic free energy pro-
file F (x) with high curvatures at both x = xts and xb,
x‡(f )/x‡ = 1 − f /x‡ × (F (cid:48)(cid:48)(xts)−1 + F (cid:48)(cid:48)(xb)−1) → 1
[5]. When free energy profile is associated with a brittle
barrier, Bell's formula can be used to extract the feature
(MFPT), k−1(f ) = D−1(cid:82) b
(cid:105)
(cid:104)−(cid:82) f
a dyeβ(U (y)−f·y)(cid:82) y
FIG. 1. Rupture characteristics obtained numerically using
a potential with two barriers at constant loading rates. (a)
U (x) (magenta) and U (x) − f · x (cyan) with A = 5 pN · nm
and f = 50 pN. Reflecting and absorbing boundary condi-
tions are set at x = a and x = b, respectively. (b) Rupture
force distributions, P (f ) = k(f )/rf · exp
0 df(cid:48)k(f(cid:48))/rf
,
at varying rf were computed by using mean first passage time
a dze−β(U (z)−f·z),
starting from the first bound state at a(=0 nm) to reach an
absorbing boundary at b(=5 nm). MFPT expression is valid
in the force regime where stationary flux approximation holds
[4]. The length was scaled by nm, and D = 1.0 × 107 nm2/s
was used for the diffusion constant. (c) [f∗, log rf ] plots at
three A values. Fits of [f∗, log rf ] to Eq.5 yield ν (cid:28) 0.5 for
all A values (ν = 0.064, 0.075, 0.046 for A = 4, 5, 6 pN · nm,
respectively). In this case, the data should be divided into
two regions and analyzed by the two linear fits as depicted
using green lines on the curve with A = 6pN · nm. (d) load-
ing rate dependent x‡(rf )(= xts − xb), extracted from the
slope of plot at each rf in (c) with A = 5 pN · nm, shows a
sharp decrease from ∼3 nm to < 1 nm around rf ≈ (e−3− e0)
pN/s.
ment of the inner transition state (see Fig.5C in Ref.(5)).
Although it is straightforward to interpret that the two
discrete slopes in Fig.1c (or the precipitous transition of
x‡ in Fig.1d) are due to crossing two barriers since the
underlying potential is given in Fig.1a, it is nontrivial to
solve the inverse problem of unambiguously determining
from [f∗, log rf ] plots and decide whether the underly-
ing free energy profile has a single barrier with a moving
transition state as rf increases or multiple barriers.
To establish a criterion for ascertaining whether the
energy landscapes for forced-ligand rupture from biotin
and LFA-1 have multiple barriers, we study the range of
applicability of DFS formalism based on a model poten-
-f!xx=ax=b(a)x2++~1.3nmx1++(outer)~3.3nmx2++x1++(inner)~0.9nm0306090120f [pN]00.10.2P(f)10-2 pN/s10-1pN/s1 pN/s10 pN/s102pN/s103pN/s-20-1001020log rf (pN/s)050100150f* (pN)A=4pN/nm5A=5pN/nm5A=6pN/nm5Eq.(5) fit-10-50510log rf (pN/s)012345x (nm)++0306090120f [pN]00.10.2P(f)10-2 pN/s10-1pN/s1 pN/s10 pN/s102pN/s103pN/s-20-1001020log rf (pN/s)050100150f* (pN)A=4pN/nm5A=5pN/nm5A=6pN/nm5Eq.(5) fit-10-50510log rf (pN/s)012345x (nm)++(b)(c)(d)0306090120f [pN]00.10.2P(f)10-2 pN/s10-1pN/s1 pN/s10 pN/s102pN/s103pN/s-20-1001020log rf (pN/s)050100150f* (pN)A=4pN.nmA=5pN.nmA=6pN.nmEq.(5) fit-10-50510log rf (pN/s)012345x (nm)++n=2n=6n=20plastic, ductile (soft)brittle (hard)............G(x)n=1n=5n=19............3
of the underlying 1D profile from DFS data [5].
For general n, the Kramers rate equation based on the
Eq.1 under tension can be derived as:
k(ε) = κεα(n) exp (−βG‡ε(n+1)/n)
(2)
where κ is the prefactor in Kramers theory and α(n) =
χ(1−1/n) with χ = 1, 2 for odd and even n, respectively.
For a given k(f ), the most probable unbinding force is
determined by dP (f )/dff =f∗ = 0, resulting in a general
equation for f∗:
k(cid:48)(f∗) =
[k(f∗)]2
(3)
1
rf
which leads to
n+1
n =
ε
−1
βG‡ log
(cid:20) rf x‡
κkBT
ε1/n−α(n)
(cid:18)
1 − 1
βG‡
(cid:19)(cid:21)
.
nα(n)
n + 1
ε− n+1
n
(4)
Under the typical condition that rupture occurs by ther-
mal activation, i.e., f (cid:28) fc(ε ≈ 1) and βG‡ (cid:29) 1, the
most probable unbinding force is approximated as:
f∗ ≈ fc
1 −
− kBT
G‡ log
rf x‡
κkBT
(5)
(cid:20)
(cid:18)
(cid:19)ν(cid:21)
where ν = n
n+1 . In deriving Eq.5 using G(x), the large
force ε(= 1 − f /fc) (cid:28) 1 or fast loading condition, an
assumption made in obtaining the mean unbinding force
expression similar to Eq.5 [6, 12, 13], is not needed. Only
the shape of the energy potential matters in deriving Eq.5
from Eq.1. The DFS data will have a larger curvature for
smaller n, namely when the energy landscape associated
with a protein complex is more ductile (Fig. 2). Because
n = 1 (harmonic cusp), 2 (linear cubic), . . ., ∞ (Bell), ν
must satisfy the bound,
1/2 ≤ ν ≤ 1
(6)
for an arbitrary 1D profile that suffices to describe rup-
ture kinetics.
For forced-rupture of biotin-ligand complex, fits to the
entire range of the data using Eq.5 give ν in the disal-
lowed range; ν = 0.40 (biotin-streptavidin) and ν = 0.070
(biotin-avidin) (see Fig.3a). Even in biotin-streptavidin
case, the parameters extracted from the fits with ν = 0.40
and ν = 0.5 (fixed) are comparable; however, the fit with
ν = 0.40 is superior yielding both smaller relative error
and reduced chi-square value, χ2
red, than with ν = 0.5,
the lower bound of Eq. 6, that gives the maximal curva-
ture in the single-barrier picture (see Fig.3(a) and Fig.S2
in the SI). For both biotin-ligand complexes, our crite-
rion consistently suggests that the unbinding landscapes
for the complexes involve more than one barrier, as was
FIG. 3. Analysis of DFS data with large curvatures. (a) The
data obtained using biomembrane force probe (BFP) with
force constant in the range 0.1-3 pN/nm [14] were fitted to
Eq.5 (solid lines) with ν=0.40 for biotin-streptavidin (circle)
and ν=0.070 for biotin-avidin (triangle). The x‡(rf ) at each
rf is calculated on the right using the slope of four succes-
sive data points of [f∗, log rf ] plot. Analyses of data using
restricted ν values (ν = 0.5 fit is in dashed line in Fig.3a) are
in the SI (b) Analysis of DFS data of LFA-1 and its ligands,
ICAM-1 and ICAM-2 in Ref.
[16]. The fits in log-log scale
are shown on the right. In all cases, ν < 1/2 suggests that
for these complexes as well the underlying free energy profiles
must contain at least two barriers; thus multi-state fits are
required by dividing the DFS data into multiple regions as
was already surmised in [16].
[14]. Next, we analyzed
emphasized by Merkel et al.
the extensive data on LFA-1 expressed in Jurkat T cells
whose binding affinity to ICAM-1 and ICAM-2 can be
enhanced by treating the cells with phorbol myristate ac-
etate (PMA) and the divalent counterion, Mg2+. Under
all conditions the exponents that best fit the DFS data
are ν < 0.5 (Fig.3b). As originally argued by entirely
different method [16] rupture of ICAM-1 and ICAM-2
from LFA-1 is best described using a free energy profile
with at least two barriers. Taken together we arrive at a
consistent conclusion that ν < 0.5 implies that the under-
lying free energy landscape in protein-ligand complexes
has multiple barriers.
-5051015log rf [pN/s]050100150200f* [pN](cid:105)=0.40 ((cid:114)red2=0.120)(cid:105)=0.50 ((cid:114)red2=0.784)(cid:105)=0.070 ((cid:114)red2=8.48)-2024681012log rf [pN/s]012345x [nm]biotin-streptavidinbiotin-avidin++246810120100200300400f* (pN)resting ((cid:105)=0.076)PMA ((cid:105)=0.063)Mg2+ ((cid:105)=0.046)24681012log rf050100150200f* (pN)resting ((cid:105)=0.031)PMA ((cid:105)=0.029)Mg2+ ((cid:105)=0.14)-5-4-3-2-10-0.5-0.4-0.3-0.2-0.100.1log1!f*fc"#$%&'log!kBTG*logrfx*"kBT#$%&'(log1−f*/fc()log−kBTG‡logrfx‡κkBT⎛⎝⎜⎞⎠⎟246810120100200300400f* (pN)resting ((cid:105)=0.076)PMA ((cid:105)=0.063)Mg2+ ((cid:105)=0.046)LFA-1:ICAM124681012log rf050100150200f* (pN)resting ((cid:105)=0.031)PMA ((cid:105)=0.029)Mg2+ ((cid:105)=0.14)LFA-1:ICAM2246810120100200300400f* (pN)resting ((cid:105)=0.076)PMA ((cid:105)=0.063)Mg2+ ((cid:105)=0.046)LFA-1:ICAM124681012log rf050100150200f* (pN)resting ((cid:105)=0.031)PMA ((cid:105)=0.029)Mg2+ ((cid:105)=0.14)LFA-1:ICAM2(a)(b)Mathematically the inequality (Eq.6) is not strict be-
cause it is possible to construct 1D profiles with ν < 0.5
for which [f∗, log rf ] plots exhibit curvature like those ob-
served in experiments. However, such free energy profiles
are physically pathological with non-existing first deriva-
tives in the vicinity of the bound complex and fits to the
data give manifestly unrealistic parameters (see Support-
ing Information for detailed calculations and analysis).
For the physical free energy profiles Eq.6 is rigorously
satisfied. In addition, there is no compelling reason to
choose a special n value even if 1D profile is deemed ad-
equate, and thus ν ought to be treated as a parameter.
Although Ref.
[11] used ν as a free parameter, the va-
lidity range for ν was not discussed.
If a global fit of
[f∗, log rf ] data using Eq.5 yields ν < 0.5 and the effect
of probe stiffness [18] is absent in the DFS data (see be-
low), we can conclude that a single barrier description of
the energy landscape is inadequate.
1
ef f = k−1
In principle curvature in the DFS data could also arise
due to probe stiffness. Simple procedure of tiling free en-
ergy profile by the amount −f · x under tension is widely
used in analyzing single molecule force experiment. How-
ever, more rigorous formulation for the effective free en-
ergy profile under load using a transducer with stiffness
2 kef f (x − Xtr)2
k should read Gtot(x, Xtr) = G(x) + 1
where x is the position of the end of molecule, Xtr
is the position of transducer, and kef f is the effective
stiffness of molecular construct combining the trans-
ducer and the complex (k−1
In fact,
2 kef f (x − Xtr)2 = −f · x + 1
tr with
f = kef f Xtr. Therefore, the effective free energy for
the complex under tension should be written in general
as Gef f (x) = G(x) − [f − 1
2 kef f x] · x [19]. As long as
f (cid:29) 1
2 kef f x (or Xtr (cid:29) x/2) especially when kef f is
small as in optical tweezers or BFP, one can approx-
imate Gef f (x) ≈ G(x) − f · x. Otherwise, rebinding
from transient capture well created by a large probe stiff-
ness at near-equilibrium loading condition could give rise
to the nonlinearity in the DFS data [18]. Therefore,
the rupture force being measured should be replaced by
f∗ → f∗ − 1
most probable force measured by using a transducer with
high stiffness such as AFM could be approximated as,
2 kef f x‡ε1/n, and at low forces (f (cid:28) fc) the
tr + k−1
m ).
2 kef f X 2
2 kef f x2 + 1
f∗ ≈ 1
2
(cid:124)
kef f x‡
(cid:123)(cid:122)
(cid:125)
+ fc
(cid:124)
=fpl
rf x‡
κkBT
− kBT
G‡ log
(cid:123)(cid:122)
=fDFS
(cid:18)
(cid:20)
1 −
(cid:19)ν(cid:21)
(cid:125)
.
(7)
The effect of probe stiffness manifests itself as a non-
vanishing plateau force, which could be as large as fpl ≈
(10 − 100) pN when kef f ≈ 100 pN/nm and x‡ = 0.1 − 1
nm, even when rf is small enough that fDFS = 0 [18].
The biotin-ligand complexes data preclude this possi-
bility because the probe stiffness of BFP kef f = 0.01−0.3
4
pN/nm [14], which is 1-2 orders of magnitude smaller
than the kef f value discussed in the literature [20, 21].
The value of kef f is 0.5 − 2.0 pN/nm in the experiments
involving LFA-1 [16]. Even the largest estimated x‡ value
for the outmost barrier (x‡ ≈ 3 nm) only yields fpl < 1
pN. Furthermore, if the nonlinear curvature of DFS data
is suspected to be due to the stiffness effect, this ought
to be discerned from the curvature due to multiple barri-
ers by producing DFS data at a reduced probe stiffness.
The curvature due to multiple barrier should persist in
the DFS data even with a small kef f . Thus, the cur-
vature in the data in [14] can only be attributed to the
presence of multiple barriers.
The condition (Eq.6) for single-barrier based 1D theo-
ries for DFS [7 -- 13] provides a guideline to judge whether
the curvature in DFS data is due to multiple barriers or
single barrier with a ductile transition state. Our work,
which does not consider complications due to various
multidimensional landscape scenarios [22, 23], shows that
the extracted parameters from the data for the protein
complexes with ligands using 1D profile with multiple
barriers are physically reasonable [14, 16]. Additional
justification for the use of such energy landscapes can
only be made by studying the structures of the protein
complexes in detail.
Acknowledgements: This work was funded by Na-
tional Research Foundation of Korea (2010-0000602)
(C.H.) and National Institutes of Health (GM089685)
(D.T.). We thank Korea Institute for Advanced Study
for providing computing resources.
[1] D. Thirumalai, E. P. O'Brien, G. Morrison, and
C. Hyeon, Annu. Rev. Biophys. 39, 159 (2010).
[2] A. Borgia, P. Williams, and J. Clarke, Annu. Rev.
Biochem. 77, 101 (2008), ISSN 0066-4154.
[3] S. Garcia-Manyesa, L. Dougana, C. L. Badillaa, J. Brujic,
and J. M. Fernandez, Proc. Natl. Acad. Sci. USA 106,
10534?10539 (2009).
[4] E. Evans and K. Ritchie, Biophys. J. 72, 1541 (1997).
[5] C. Hyeon and D. Thirumalai, J. Phys.: Condens. Matter
19, 113101 (2007).
[6] A. Garg, Phys. Rev. B. 51, 592 (1995).
[7] G. Hummer and A. Szabo, Biophys. J. 85, 5 (2003).
[8] O. K. Dudko, A. E. Filippov, J. Klafter, and M. Urbakh,
Proc. Natl. Acad. Sci. USA 100, 11378 (2003).
[9] Y. Sheng, S. Jiang, and H. Tsao, J. Chem. Phys. 123,
091102 (2005).
[10] O. K. Dudko, G. Hummer, and A. Szabo, Phys. Rev.
Lett. 96, 108101 (2006).
[11] O. K. Dudko, J. Math´e, A. Szabo, A. Meller, and
G. Hummer, Biophys. J. 92, 4188 (2007).
[12] H. J. Lin, H. Y. Chen, Y. J. Sheng, and H. K. Tsao,
Phys. Rev. Lett. 98, 088304 (2007).
[13] R. W. Friddle, Phys. Rev. Lett. 100, 138302 (2008).
5
[14] R. Merkel, P. Nassoy, A. Leung, K. Ritchie, and E. Evans,
Nature 397, 50 (1999).
[15] M. Schlierf and M. Rief, Biophys. J. 90, L33 (2006).
[16] E. P. Wojcikiewicz, M. H. Abdulreda, X. Zhang, and
V. T. Moy, Biomacromolecules 7, 3188 (2006).
[17] C. Hyeon and D. Thirumalai, Biophys. J. 90, 3410
(2006).
[18] E. Evans, Annu. Rev. Biophys. Biomol. Struct. 30, 105
(2001).
[19] E. B. Walton, S. Lee, and K. J. Van Vliet, Biophys. J.
94, 2621 (2008).
[20] R. Friddle, P. Podsiadlo, A. Artyukhin, and A. Noy, J.
Phys. Chem. C 112, 4986 (2008).
[21] Z. Tshiprut, J. Klafter, and M. Urbakh, Biophys. J. 95,
L42 (2008).
[22] B. T. Marshall, M. Long, J. W. Piper, T. Yago, R. P.
McEver, and C. Zhu, Nature 423, 190 (2003).
[23] V. Barsegov and D. Thirumalai, Proc. Natl. Acad. Sci.
USA 102, 1835 (2005).
SUPPORTING INFORMATION
DFS theory for a free energy profile with non-
integer n.
It could be argued that the inequality
1/2 ≤ ν ≤ 1 (valid rigorously for integer n) that accounts
for the curvature of DFS data is not mathematically re-
quired. Here we show that it is possible to construct 1D
free energy profiles for which ν is clearly less than 0.5.
Indeed, these free energy profiles can even have nearly
vanishing ν. However, such profiles are unphysical be-
cause near the bound state they have incorrect curva-
tures compared to the physical profiles discussed in the
text and in the references cited therein. More impor-
tantly, the first derivatives of these free energy profiles,
which yield ν < 0.5 do not exist near the bound state
i.e, they have a singularity. For these and other reasons
(see below) we reject these free energy profiles as plausi-
ble models for explaining the curvatures in the observed
[f∗, log rf ] plots in a number of protein complexes dis-
cussed in the main text, which have all been explained
using a two barrier model.
A free energy profile with 0 < n < 1 (see Eq.1) can
n+1 . To explore this
lead to 0 < ν < 1/2 since ν = n
possibility, we consider a free energy profile,
G(x) = σax1/σ − bx
(S1)
with σ > 1. Here we may regard n = 1/σ, and hence
n < 1. The term −bx is required for the potential to
have a barrier at a finite value of x‡, the location of the
transition state (TS). The TS location and the associated
barrier height are
(cid:16) a
(cid:17) σ
σ−1
x‡ =
G‡ = (σ − 1)a
b
(cid:16) a
(cid:17) 1
σ−1
b
.
(S2)
Under tension f the potential becomes Gef f = G(x) −
f x = σax1/σ − (b + f )x. The f -dependent TS location
and the barrier height are
σ−1
(cid:19) σ
(cid:19) 1
σ−1
(cid:18) b
(cid:18) b
b + f
b + f
G‡/x‡
G‡/x‡
σ−1
G‡/x‡
σ−1 + f
σ−1
G‡/x‡
σ−1 + f
σ−1
σ
1
σ−1
=
=
x‡(f )
x‡ =
G‡(f )
G‡ =
σ
1−σ
= η
= η
1
1−σ .
(S3)
where η ≡ (1 + f /f1/σ) with f1/σ ≡ 1
G‡
x‡ . Therefore,
σ−1
one can rewrite Eq.S1 (G(x)) and an effective free energy
(Gef f (x)) under tension as
G(x) =
Gef f (x) =
σG‡
σ − 1
σG‡
σ − 1
(cid:16) x
(cid:16) x
x‡
x‡
(cid:17)1/σ − G‡
(cid:17)1/σ − (
σ − 1
G‡
σ − 1
(cid:16) x
(cid:17)
,
x‡
+ f x‡)
6
(cid:17)
(cid:16) x
x‡
. (S4)
Given Gef f (x), it is possible to obtain the mean first
passage time expression corresponding to the lifetime of
the complex that can be measured in single molecule ex-
periments. It is given by,
(cid:90) x
(cid:90) ∞
ef f (0)y(cid:17)−1(cid:113) G(cid:48)(cid:48)
dxeβGef f (x)
0
0
approximation,
ef f (x‡(f ))
2πkB T
k(f )−1 =
1
D
(cid:16)(cid:82)
The
saddle-point
bound dye−βG(cid:48)
D
Eq.A3, yields Kramers' equation for the escape rate:
k(f )
≈
e−βG‡(f ) with
dye−βGef f (y).
(S5)
(cid:104)−βG‡η
(cid:105)
1
1−σ
k(f ) = κηα(σ) exp
(S6)
(cid:16)(cid:82)
ef f (0)y(cid:17)−1
with α(σ) ≡ 2σ−1
2(σ−1) . Here, note that the κ, de-
fined as the prefactor, contains the singular integral
bound dye−βG(cid:48)
. By using the relationship
k(cid:48)(f∗) = [k(f∗)]2/rf to derive the most probable force,
we get
(cid:20) (σ − 1)α(σ)
(cid:21)
ηα(σ)+1βG‡ + η
σ
1−σ
.
(S7)
η
1
1−σ = − 1
βG‡ log
rf x‡
κkBT
By assuming βG‡ (cid:29) 1 and f (cid:28) f1/σ, we can simplify
the above equation into
η
1
1−σ ≈ − 1
βG‡ log
rf x‡
κkBT
.
(S8)
Therefore, the most probable force (f∗) for the fractional
potential (Eq.S1) is
(cid:34)(cid:18)
(cid:34)(cid:18)
f∗ ≈ f1/σ
= f1/σ
− 1
βG‡ log
− 1
βG‡ log
rf x‡
κkBT
rf x‡
κkBT
(cid:35)
(cid:19)1−σ − 1
(cid:19)2−1/ν − 1
(cid:35)
,
(S9)
where σ = 1−ν
ν was employed in the last line.
There are a few important comments about Eq.S9 that
need to be made. (i) Note that the form of Eq.S9 is very
different from Eq.5. The difference is related to the afore-
mentioned difficulties associated with Gef f (x). Never-
theless, it is possible mathematically to construct model
free energy profiles, without regard to physical consid-
erations, for which [f∗, log rf ] plot has curvature that is
reminiscent of what is observed in experiments. (ii) Al-
though Eq.S9 can be used to obtain excellent fits to DFS
7
FIG. S1. (a) The blue curve is the bare (f = 0) free energy profile of the form given in Eq. (S1) and the green curve is
the tilted form of G(x) in the presence of force. By fitting the numerically computed (black circles) f∗ as a function of rf to
Eq.5 (red curve) we obtain the parameters shown below. Although the features of original potential G(x) = 100x1/3 − 10x are
reasonably recovered (σ is larger than the value in G(x)) by using Eq.5, the extracted value of κ is unrealistically large. (b)
Eq.5 was used to fit the DFS data of biotin-strepavidin (circles) and biotin-avidin (triangles). Although the quality of fit is
excellent, the unrealistically large value of κ, due to the singularity of the hypothesized fractional potential at x = 0, suggests
that the potential with a fractional power should not be used for the analysis. In fact the κ values are comparable to or much
greater than the TST estimate kBT /h (≈ 6.2 × 1012 s−1), which of course makes no sense. Hence, we can rule out the free
energy profiles of the form given in Eq.S1 to analyze DFS data on protein complexes.
data on protein complexes, it turns out that the extracted
value of κ is extremely large and are clearly unphysical
(see Fig.S1). The unphysical values are a consequence of
the singularity of G(x) (or Gef f (x)) at x = 0. We con-
clude that the free energy profiles with a fractional power
of n, which mathematically creates singularity at bound
state, is not suitable to be used to analyze experimental
data. Thus, the bound 1/2 < ν < 1 in Eq.6 must hold
for physical 1D free energy profiles with a single barrier.
-20-1001020log rf020406080100f*05101520x (nm)050100150200G(x) (pN.nm)G(x)=100x1/3-10xGeff(x)=G(x)-10xf*=8.54−149.2(logrf−36.3)⎛⎝⎜⎞⎠⎟1−4.17−1⎡⎣⎢⎢⎤⎦⎥⎥G‡=49.2kBT, x‡=7.51 nm, κ=3.0×1015s-1, σ=4.17-5051015log rf [pN/s]050100150200f* [pN] f*=422.1((-(log rf-112.6)/19.1)1-1.13-1) G*=19.1kT, x*=1.42nm, (cid:103)=2.68×1048s f*=11.15((-(log rf-27.5)/31.1)1-5.29-1) G*=31.1kT, x*=2.70nm, (cid:103)=5.59×1011f*=422.1−119.1(logrf−112.6)⎛⎝⎜⎞⎠⎟1−1.13−1⎡⎣⎢⎢⎤⎦⎥⎥G‡=19.1kBT, x‡=1.42 nm, κ=2.68×1048s-1, σ=1.13f*=11.15−131.1(logrf−27.5)⎛⎝⎜⎞⎠⎟1−5.29−1⎡⎣⎢⎢⎤⎦⎥⎥G‡=31.1kBT, x‡=2.70 nm, κ=5.59×1011s-1, σ=5.29(a)(b)8
exp/f∗
f it − f∗
FIG. S2. Analysis of DFS data using Eq.5 for (a) biotin-streptavidin and (b) biotin-avidin. For each ν, the fits, residuals
(f∗
exp × 100), and extracted parameters were summarized in the table on the right. We can draw some general
conclusions from the fits. For the biotin-streptavidin complex, fit with ν = 0.397 produces the smallest errors although at high
loading rates the relative errors for ν = 0.397 and ν = 0.5 are comparable. There are variations in other parameters (x‡, G‡,
and κ) for all ν. For the biotin-avidin complex the situation is far worse. In particular, the relative errors in the fits are large
even when ν is varied. Similarly, the parameters extracted from the fits are not totally consistent. Taken together, the fits
using a 1D free energy profile with a single barrier is not appropriate to describe the rupture kinetics of these two complexes.
(a)ν x‡ (nm)G‡ (kT) κ (s-1)0.3970.90015.661.91×1040.1201/20.71914.641.22×1040.7842/30.54713.569.24×1035.743 x‡ (nm)κ exp(-G‡/kT) (s-1)κ exp(-G‡/kT) (s-1)10.3960.0160.01615.32Double barrier fit with ν=1Dynamics over inner barrierDynamics over outer barrier 0.1340.5720.2490500.249050χred2050100150200f* [pN](cid:105)=0.4(cid:105)=1/2(cid:105)=2/3(cid:105)=12 barrier fit-5051015log rf [pN/s]051015202530rel. error (%)(b)ν x‡(nm)G‡ (kT)κ (s-1)0.070.90613.32.70×1048.481/20.58711.951.03×10421.282/30.48811.748.54×10333.67 x‡(nm)κ exp(-G‡/kT) (s-1)κ exp(-G‡/kT) (s-1)10.3260.0460.04670.48Triple barrier fit with ν=1Dynamics over inner barrierDynamics over middle barrierDynamics over outer barrier 0.1120.3213.890.0220.0658700.0220.065870χred2050100150200f* [pN](cid:105)=0.070(cid:105)=1/2(cid:105)=2/3(cid:105)=13 barrier fit-5051015log rf [pN/s]050100150200rel. error (%) |
1006.4431 | 1 | 1006 | 2010-06-23T07:10:24 | On the feasibility of nanocrystal imaging using intense and ultrashort 1.5 {\AA} X-ray pulses | [
"physics.bio-ph",
"physics.comp-ph"
] | Structural studies of biological macromolecules are severely limited by radiation damage. Traditional crystallography curbs the effects of damage by spreading damage over many copies of the molecule of interest. X-ray lasers, such as the recently built LINAC Coherent Light Source (LCLS), offer an additional opportunity for limiting damage by out-running damage processes with ultrashort and very intense X-ray pulses. Such pulses may allow the imaging of single molecules, clusters or nanoparticles, but coherent flash imaging will also open up new avenues for structural studies on nano- and micro-crystalline substances. This paper addresses the theoretical potentials and limitations of nanocrystallography with extremely intense coherent X-ray pulses. We use urea nanocrystals as a model for generic biological substances and simulate primary and secondary ionization dynamics in the crystalline sample. Our results establish conditions for ultrafast nanocrystallography diffraction experiments as a function of fluence and pulse duration. | physics.bio-ph | physics |
On the feasibility of nanocrystal imaging using intense and ultrashort 1.5 A X-ray
pulses
C. Caleman1,2, G. Huldt3, F. R. N. C. Maia3, C. Ortiz4, F. G. Parak1,
J. Hajdu3, D. van der Spoel3, H. N. Chapman2,5 and N. Timneanu3∗
1Physik Department E17, Technische Universitat Munchen,
James-Franck-Strasse, DE-85748 Garching, Germany
2Center for Free-Electron Laser Science, DESY, Notkestrasse 85, DE-22607 Hamburg, Germany
3Department of Cell and Molecular Biology, Biomedical Centre,
Box 596, Uppsala University, SE-75 124 Uppsala, Sweden
4Institut fur Theoretische Physik, Goethe-Universitat Frankfurt,
Max-von-Laue-Strasse 1, DE-60438 Frankfurt am Main, Germany
5 University of Hamburg, Luruper Chaussee 149, DE-22761 Hamburg, Germany
Structural studies of biological macromolecules
are severely limited by radiation damage. Tradi-
tional crystallography curbs the effects of dam-
age by spreading damage over many copies of
the molecule of interest. X-ray lasers, such as
the recently built LINAC Coherent Light Source
(LCLS) [1], offer an additional opportunity for
limiting damage by out-running damage pro-
cesses with ultrashort and very intense X-ray
pulses. Such pulses may allow the imaging of sin-
gle molecules, clusters or nanoparticles, but co-
herent flash imaging will also open up new av-
enues for structural studies on nano- and micro-
crystalline substances. This paper addresses the
theoretical potentials and limitations of nanocrys-
tallography with extremely intense coherent X-
ray pulses. We use urea nanocrystals as a model
for generic biological substances and simulate pri-
mary and secondary ionization dynamics in the
crystalline sample. Our results establish condi-
tions for ultrafast nanocrystallography diffraction
experiments as a function of fluence and pulse du-
ration.
Any sample exposed to an intense X-ray pulse will
be ionized and extensive ionization destroys the sample.
The time scale on which this process occurs is critical
for obtaining an interpretable diffraction pattern that
conveys an atomic structure of the sample.
In princi-
ple, the X-ray pulse must be short enough for the en-
tire pulse to pass through the sample before major disar-
rangement of atomic and electronic configurations takes
place. The ionizations due to direct photoabsorption and
subsequent secondary processes affect the ability to get
useful structural information from the diffraction pat-
tern in three ways: (i) Ionization decreases the elastic
X-ray scattering power of atoms and induces consid-
erable changes in diffracted intensities due to ionization
stochasticity. (ii) Removal of electrons from atoms leaves
behind positively charged ions that repel each other due
∗E-mail: [email protected]
FIG. 1: Comparison of crystal size and the modeled
size of secondary electron cascades. (a) The unit cell of
a urea crystal contains light elements abundant in proteins:
carbon (depicted as green), nitrogen (blue), oxygen (red) and
hydrogen (white). (b) A urea nanocrystal of 200 nm would
contain about 50 million unit cells. A protein nanocrystal
of similar size would contain about 20,000 unit cells (using
lysozyme as an example). (c) The overall dimensions of sim-
ulated electron clouds produced during the thermalization of
a single 0.4 keV Auger electron ejected from a nitrogen atom
(top) and a single 8 keV photoelectron (bottom) inside a large
urea crystal after 50 femtoseconds. Similar cascade sizes are
produced in protein crystals, during an X-ray diffraction ex-
periment. The total number of ionizations was 18 in the Auger
cascade, and 390 in the photoelectron cascade at 50 fs after
the emission of primary electrons. At this point, the radius of
gyration of the photoelectron cloud reached 300 nm, and that
of the Auger electron cloud 8 nm. The photoelectron cascade
is bigger than a typical nanocrystal under consideration here.
to Coulomb forces, leading to the destruction of the struc-
ture. (iii) Free electrons either leave the sample, if their
energy is high enough, or remain in the sample as a back-
ground electron gas, in which case they will be a source
of noise in the diffraction image.
2
FIG. 2: Evolution of secondary ionization cascades in a urea crystal over time. (a) Number of secondary ionizations
produced by a photoelectron of 8 keV and by Auger electrons (impact energies: 250 eV for carbon, 400 eV for nitrogen,
500 eV for oxygen). (a) Spatial evolution of the electron cloud from a single photoelectron of 8 keV energy in an infinitely
large urea crystal depicted through the radial electron density as a function of time and radial distance from the point of
incidence. The cascade includes the primary electron and its secondary electrons. At any given time, the 4π integration over
the radius in the 3D volume gives the total number of electrons, assuming spherical symmetry when the cascades are added
stochastically. (c) Secondary electron cloud from a single Auger electron (nitrogen). The cascade includes the primary electron
and its secondary electrons. The thermalization of electrons from oxygen and carbon has similar features. Black lines in (b)
and (c) show the radii of gyration (Equation 1) of the electron clouds. In the first femtosecond the electron clouds are highly
anisotropic. After 20 fs no more secondary ionizations will occur in the photoelectron cascade (5 fs for Auger cascades). Figures
show averaging over 1000 simulations on an infinitely large urea crystal.
Thermalization of trapped electrons leads to additional
ionizations through cascade processes. The probability
of trapping, and the size of the resulting secondary elec-
tron cascades, depends on sample size and X-ray energy
(among others). Photoelectrons released by X-rays of
1.5 A wavelength are fast (53 nm/fs) and can escape from
small samples such as "nanosized" crystals [2, 3] early
in an exposure (Figure 1). In contrast, Auger electrons
from carbon atoms are slow (9.5 nm/fs) and cause sec-
ondary ionization even in a single protein molecule [4, 5].
For small samples (diameter<10 nm), the explosion is
dominated by the repulsion of positive ions left behind
by electrons leaving the sample [4, 6].
In big samples
(diameter>500 nm), most electrons will be trapped sim-
ply because they lose energy before reaching the surface.
Trapped electrons increase the temperature of the sample
through collisional processes, while slowing the explosion
by partially screening the positive charges and creating a
net neutral core. Predictions point to a transition from
Coulomb explosion to a hydrodynamic expansion with in-
creasing sample size. A positively charged surface layer is
formed, destroying the sample from outside towards the
center. The expansion is driven by thermal processes as
the electron pressure grows [7]. Hence, crystalline diffrac-
tion and useful structural information might still be ob-
tained from the center of the crystal.
Descriptions of electron impact ionization and sec-
ondary ionization cascades exist for different materi-
als [5, 8, 9]. Dynamics of photoelectrons in protein crys-
tals have been investigated earlier [2, 3], without consid-
eration to Auger emission or secondary ionization cas-
cades. These predictions suggest that radiation damage
can be limited by reducing the crystal size. The present
work steps beyond these studies and gives an integrated
description of photo-emission, Auger emission and cas-
cade processes during exposure of a biological nanocrys-
tal to an ultrashort and intense X-ray pulse, to determine
the feasibility of nanocrystal imaging and improvement
in resolution achievable with shorter pulses. Our find-
ings are summarized in Figures 1-4 and the methodology
is described in Methods and Supplementary.
For light elements, a single photo-ionization releases
electrons at two distinctly different energies (Figure 1).
The photoelectron energy corresponds to the difference
between the photon energy and the K-shell binding en-
ergy, while Auger electrons carry kinetic energy depen-
dent on atom type (250 eV for carbon). The average
time for the first collisional ionization scales with the
primary electron energy (Figure 2a). The electron cloud
initiated by a photoelectron thermalizes slower than elec-
trons in Auger cascades, since energetic electrons travel
further between scattering events in the crystal due to
their lower interaction cross section (Figures 2b,c). At
the same time, the cloud generated by a photoelectron is
around four orders of magnitude larger in volume than
the Auger electron induced cloud. After thermalization,
the electron clouds keep expanding through diffusion, fol-
lowing a random walk pattern. Figures 2b,c show that
the radius of gyration (Equation 1) at these impact en-
ergies describes well the spatial extent of the electron
clouds.
In a sample that is small compared to the size of the
X-ray beam or the photon absorption length, photoion-
ization events will occur with equal probability through-
out the entire sample. At 8.3 keV photon energy, a sin-
gle photoelectron will liberate about 400 electrons before
reaching thermal equilibrium (Figure 2a). The electron
gas will have high temperature due to the high photon en-
3
FIG. 3: Degradation of the detected signal as a result of radiation damage. (a) Decay of Bragg peaks during exposure
to an X-ray pulse in a urea crystal. Pulse length: 10 fs (FWHM) centered at t=0, at 1.5 A wavelength. The pulse intensity
(1.5×1011 photons in a focal spot of 1 µm diameter FWHM) is such that atoms are ionized once in average when 99.5% of
the intensity of the pulse passes through the sample. Peak intensities at different resolutions are represented by the (hkl)
reflections and are normalized to one, based on intensities from undamaged crystals. The (330) reflection corresponds to 1.2 A
resolution and has a pulse-integrated degradation of 50% due to ionization and atomic displacement. The dashed black line
shows the average root mean square deviations (RMSD) in atomic positions during illumination. (b,c) Contour plots for the
average ionization per atom (¯z) and the R-value as a function of the X-ray pulse length and intensity (photons/focal spot of 1
µm diameter FWHM). The R-value (Equation 3) is calculated from all the Bragg peaks up to a resolution of 1.5 A. The red
thick line corresponds to an R-value of 0.15, lower values are considered acceptable for a good reconstructable signal. The blue
thick dashed line represents the damage of 1 ionization per atom. The plot in (b) shows behavior of nanocrystals smaller than
200 nm, from which the photoelectrons escape early in an exposure, while (c) shows the behavior of crystals larger than 500
nm, when photoelectrons are completely trapped during exposure.
ergy, and the electrons will be distributed approximately
uniform throughout the sample. The free electrons will
scatter predominantly in the forward direction and con-
tribute incoherently to the background in the diffraction
pattern.
At 1.5 A wavelength, the ratio between elastically scat-
tered photons and photoionization is 1:32 for oxygen, 1:26
for nitrogen and 1:20 for carbon. Incoming photons will
primarily ionize sample and only a few will contribute
to coherent scattering. The loss of an electron will de-
crease the scattering power by 17% for a carbon atom,
14% for nitrogen and 12% for oxygen. One ionization
per atom also leads to atomic displacement and further
degradation of the scattered signal (Figure 3a).
Since a focused X-ray pulse will destroy the sample,
three-dimensional (3D) structure determination relies on
the experiment being repeatable. Rather than build-
ing up a complete X-ray diffraction data set by rotat-
ing the crystal and collecting a sequence of diffraction
images it will be necessary to scale together individual
diffraction images from many different nanocrystals, in
order to assemble a complete 3D data set [10]. A crys-
tal with 5×5×5 unit cells will produce a discrete diffrac-
tion pattern [11], and conventional X-ray analysis tech-
niques may be used for
indexing, merging and recon-
struction [10]. Furthermore, oversampling techniques for
direct phase retrieval may also be employed for a 3D
structural determination [12].
We express damage-induced errors in terms of degra-
dation of Bragg peaks, and for the entire diffraction
pattern we calculate an R-value from simulated crys-
tals exposed to X-ray pulses (Figure 3). The R-value
is a measure of the overall agreement between the crys-
tallographic model and the experimental X-ray diffrac-
tion data (Equation 3). Small molecules (such as urea)
form more ordered crystals and an R-value below 0.05
is considered a good threshold (Cambridge Structural
Database). For macromolecules, values up to 0.20 are
acceptable (Protein Data Bank) and we use the conven-
tion R< 0.15 from [11] (Figure 3). Figure 3a shows how
Bragg peaks may degrade during exposure to an X-ray
pulse, and how this influences the dependence of the R-
value with pulse parameters. The R-value is also de-
pendent on crystal size, and Figures 3b,c compare two
regimes: crystals where Auger cascades dominate versus
crystals where damage is driven by photoelectron cas-
cades. Trapping of photoelectrons in crystals larger than
500 nm leads to a steeper degradation of the signal and
constrains what pulse lengths and intensities could be
used for successful imaging.
When deciding which parameters of the X-ray laser
pulse and sample characteristics one should use (Fig-
ure 4), there is an interplay between two effects driven
mainly by photoelectrons; i) If photoelectrons escape the
sample, the total number of ionizations will be signifi-
cantly lowered. To reduce radiation damage early in the
exposure, the sample has to be smaller than the size of
the photoelectron cascade. The diffraction signal from
the crystal scales with the size of the crystal as a power
law (Equation 5). Thus reducing sample size will con-
versely require an increase in pulse intensity in order to
retain the signal at the same level (Figure 4).
ii) If the
0 0.2 0.4 0.6 0.8 1-15-10-5 0 5 10 15 0 0.02 0.04 0.06 0.08 0.1Relative intensity of Bragg peakRMSD atomic positions (nm)Time (fs)aPulseRMSD urea(110)(220)(330)101010111012 0 10 20 30 40 50 60 70 80 90 100Pulse intensity (photons/pulse)Pulse length (FWHM, fs)R=0.050.150.25z-=1z-=2z-=0.2bcrystal size < 200 nm R value = 0.151 ionization/atom10910101011 0 10 20 30 40 50 60 70 80 90 100Pulse intensity (photons/pulse)Pulse length (FWHM, fs)R=0.050.150.25z-=1z-=2z-=0.2z-=3z-=0.5ccrystal size > 500 nm R value = 0.151 ionization/atom4
than 10 fs for crystals larger than 500 nm, where as for
small crystals (< 200 nm) pulse lenghts can be as long
as 30 fs.
Ultrafast single-shot nanocrystallography fills the gap
between single molecule imaging and crystallography. It
offers the opportunity to investigate biological molecules
which are too small to provide a good signal on their
own in an X-ray laser diffraction experiment, however
they could form nanocrystals which would be too small
to investigate with conventional synchrotron radiation.
Methods
1.
Electron impact ionization.
Simulations
of the ionization cascade dynamics in crystalline urea
(CO(NH2)2) were performed using the spatial electron
dynamics program, ehole, that is a part of the gro-
macs [15] Molecular Dynamics software package. Urea
was chosen as model for a biological sample for three rea-
sons: it has a well known crystalline structure, it has an
atomic composition of biological character, and its unit
cell is small.
In earlier work [5], the inelastic electron
cross sections for urea have been derived from first princi-
ples calculations. Based on these we have calculated the
number of secondary electrons generated by an impact
electron in a urea crystal. The inelastic cross section for
electron scattering in urea is comparable in magnitude
with that for water [5]. Thus, urea crystals are a good
model for protein nanocrystals, known to contain 30%-
60% water. We refer to [5, 9, 16] and the supplementary
material for further details of these calculations and how
the model compares with experiments on diamond [17].
Considering mi to be the mass of electron i and ri the
position of electron i with respect to the center of mass of
all free electrons, the radius of gyration, used in Figure 2,
is defined as
(cid:18)(cid:80)
(cid:80)
i ri(t)2mi
i mi
(cid:19)1/2
Rg(t) =
.
(1)
2. Electron thermalization during the pulse. We
assume that the X-ray pulse can be described by a Gaus-
sian centered at time t0 = 0 and will consider the incom-
ing X-ray photons to be unpolarized and have a wave-
length of 1.5 A. Following this pulse, several primary
ionizations are treated -- the photoelectric effect result-
ing in an ejection of a high energy electron (≈ 8 keV),
accompanied by an Auger effect which provides an elec-
tron of a lower energy, depending on atomic species. The
emission for these electrons is described by normalized
probability distributions: (i) the photoelectric effect is in-
stantaneous so the emission probability follows the same
Gaussian profile as the X-ray pulse, with the width w;
(ii) the probability for an Auger process to be emitted is
a convolution of a Gaussian with the exponential decay
characteristic for each individual atomic species. The ex-
ponential decays are taken with corresponding life times
FIG. 4: Photon signal as a function of unit cell and
crystal size for a perfect crystal. The integrated Bragg
peak intensity (Equation 5) is shown for the lowest resolution
where a full reflection can be recorded (Equation 4). Fully
integrated Bragg peaks can be used for indexing and averag-
ing the signal from different nanocrystal orientations in the
beam, and 10 detected photons are considered to give a good
signal-to-noise ratio. The X-ray pulse has an intensity of 1011
photons focused in a focal spot of 1 µm diameter (FWHM),
wavelength 1.5 A, beam divergence is 0.1 mrad and spectral
bandwidth is ∆λ/λ = 0.1%. The solid lines correspond to
a scattered signal of 1, 10 and 100 photons in the first fully
integrated Bragg peak, when peak degradation is not taken
into account. The signal scales with the X-ray fluence, thus
the "1" line will correspond to 10 scattered photons for the
case of 1012 incident photons, and the "100" line will corre-
spond to 10 scattered photons for the case of 1010 incident
photons. If larger bandwidth or divergence is expected, full
Bragg peaks can be recorded at lower resolutions and conse-
quently the detected signal will be higher (Equation 5). The
inset shows details for crystals with small unit cells.
pulse is very short, the photoelectric cascade will not have
time to develop to reach a large number of ionizations.
At the same time, short pulses considerably reduce the
signal degradation due to atomic displacements and ion-
ization (low R-value in Figure 3). In this case the sample
size is less important, and one can investigate any crystal
size at photon fluences that will provide enough signal.
The size could however be constrained by coherence re-
quirements due to the pulse length [13]. At extremely
short pulses, one would also need to consider the broad-
ening of Bragg peaks [14], i.e. a bandwidth effect.
The above considerations stress the importance of hav-
ing very short pulses as means for radiation damage
control, to reduce both the ionization cascades and the
A photon flux of 1012 photons per
atomic disorder.
pulse and unit area (µm2) will offer the opportunity
to investigate a wider range of crystal sizes and unit
cells sizes. For lower available intensities (1010 pho-
tons/pulse), longer pulse lengths can be accommodated
and imaging nanocrystals of small proteins with a small
unit cell, such as lysozyme, could be feasible. Our cal-
culations show that to achieve an R-value of 0.15 at a
fluence of 1011 photons/µm2 pulses have to be shorter
τ of 11.3 fs for carbon, 8.3 fs for nitrogen and 6.6 fs for
oxygen. The probability for photoionization in urea is
determined by the cross section of the atoms, which are
well known For the three atomic species that can un-
dergo an Auger process, the contribution from the atoms
C, N, and O, is weighted according to the photoioniza-
tion cross section on the respective atoms, σC, σN, σO,
and normalized to the total photoelectric cross section
for the urea molecule. The single electron ionization cas-
cades develop mainly along the direction of the incident
photon, however we consider spherical symmetry when
these are stochastically produced. Thus, the entire ion-
ization cascade following an X-ray pulse impinging on a
crystal is given by
(cid:88)
(cid:26)(cid:90)
(cid:90)
C(t) =
+
niσi
i=C,N,O
− (t(cid:48)−t0)2
2w2
N e
1
τ
N e
− (t(cid:48)−t0)2
2w2 Cphoto(t, t(cid:48))dt(cid:48)+
τ CAuger(t, t(cid:48))dt(cid:48)(cid:27)
,
e− (t−t(cid:48) )
(2)
where Cphoto(t, t(cid:48)) and CAuger(t, t(cid:48)) represent the cascade
development with time for a single electron starting from
time t(cid:48). These are obtained from MD simulations and are
represented by the ionization rate as a function of time
(Figure 2a for t(cid:48) = 0), or radii of gyration, Figures 2b,c.
3. X-ray interactions and damage quantification.
The degradation of Bragg peaks in Figure 3 has been cal-
culated from MD simulations on an urea crystal, using
gromacs with a stochastic interaction of X-ray photons
with atoms, assuming unpolarized X-rays and homoge-
neous spatial distribution of the free electrons. The sim-
ulation box was 10x10x10 unit cells of urea, with peri-
odic boundary conditions, and includes thermal motion
of atoms. The model is described in reference [11], and
in the supplementary material. The intensity of Bragg
peaks at each time step is defined by integrating around
each peak over a rectangular area centered on the Bragg
peak and with sides of length equal to 1/10 of the sepa-
ration between adjacent peaks [18]. The spectral width
∆λ/λ, beam divergence or any broadening of the Ewald
sphere are not taken into account here. The degradation
of the Bragg peak is expected to be smaller when inte-
grating through an Ewald sphere of finite thickness. To
estimate the damage induced error we make use of the
R-value (used in Figure 3), calculated up to a resolution
q from
(cid:80)
(cid:12)(cid:12)(cid:12)(cid:112)(cid:104)Ihkl(cid:105)t −(cid:112)I 0
(cid:112)(cid:104)Ihkl(cid:105)t
hkl<q
(cid:80)
hkl
(cid:12)(cid:12)(cid:12)
R(q) =
hkl<q
,
(3)
where the summation is performed over all Bragg peaks
(hkl) corresponding to scattering vectors less than q. The
intensities of the Bragg peaks I 0
hkl for the undamaged
crystal are used as reference when compared with the
time averaged intensities (cid:104)Ihkl(cid:105)t of the damaged crystal
exposed to a Gaussian-shaped X-ray pulse. The latter
5
intensities take into account the ionization dynamics and
atomic displacement as a function of time during the X-
ray pulse.
4. Minimum required signal.
In ultra-fast single-
shot experiments at X-ray lasers the crystals will be ex-
posed in random orientations, X-rays beams are expected
to have small divergence (< 1 mrad) and are highly
monochromatic (spectral bandwidth ∆λ/λ < 0.1%).
Thus, single shot diffraction patterns will contain many
partially reflected Bragg peaks. Full Bragg peaks may
be recorded on the detector at higher resolution, and
could be used for retrieving the original orientation of
the nanocrystals and for averaging the signal from simi-
lar orientation [10] (partial Bragg peaks could in princi-
ple also be used for indexing). In our estimates for the
minimum required signal for successful indexing, we con-
sider only the signal from fully integrated Bragg peaks
at the lowest resolution where these can be recorded.
For given parameters that control the thickness of the
Ewald sphere (beam divergence ∆φ, spectral bandwidth
∆λ/λ), the lowest resolution where a full Bragg peak can
be recorded (qmin) is found by comparing the thickness of
the Ewald sphere to the size of the Bragg peak (modeled
as inversely proportional to the crystal width A)
(cid:114)
1
A
∼ ∆λ
λ
λ
2
q2
min + ∆φ qmin
1 − λ2
4
q2
min .
(4)
The average number of photons scattered elastically by
a protein crystal within a Bragg peak for a given resolu-
tion qmin can be approximated by the expression for the
integrated reflected intensity [19]
IBragg(qmin) ≈ I0(λ)
λ2
1 + cos2 θ0
(cid:88)
2 sin2(θ0/2)
×λ2 A3
a6 r2
e
atoms
2
×
(5)
f 2
atom(θ0) ,
where both small beam divergence and polychromatic-
ity are accounted for. I0(λ) is the spectral intensity of
the incoming X-ray beam (number of incoming photons
per unit wavelength and unit area), integrated over the
angular density of the incident beam. The Lorentz fac-
tor λ2/(2 sin2(θ0/2)) takes into account the integration
over the thickness of the Ewald sphere encompassing the
Bragg peak, in a similar way as [19] for the stationary
case (no rotation) with divergent polychromatic radia-
tion. The polarization factor is given by (1 + cos2 θ0)/2.
Furthermore, re is the classical electron radius, λ the
wavelength, A the crystal side (cubic crystal), a the unit
cell side (cubic unit cell), fatom the atomic scattering fac-
tor, and θ0 the polar angle between the incident pulse and
the center of the Bragg peak.
It is assumed that the unit cell structure factor is con-
stant within the Bragg peak, and that adjacent Bragg
peaks do not overlap; both these approximations improve
with the ratio A/a. The squared structure factor of the
unit cell is represented by its average value at high scat-
tering angles [20]. For numerical evaluation, the unit
cell was assumed to have a density of 1/30 A−3 carbon-
equivalent atoms (corresponding to a unit cell consist-
ing of 50% non-structural water and protein with density
approximately 1.35 g/cm3), and the scattering factor of
carbon was calculated from the Cromer-Mann parame-
ters [21]. The calculation of the number of photons per
Bragg peak (Equation 5) presented in Figure 4 with the
assumption of perfect crystals with no mosaicity. It has
been shown that micron sized crystals could consist of
only a few highly ordered domains [22], thus nanocrys-
tals are unlikely to be organized with a mosaic spread.
The above approximations will break down when the
crystal size approaches the unit cell size, as the diffracted
image turns from a discrete into a continuous pattern.
6
Acknowledgments
The Swedish Research Council is acknowledged for fi-
nancial support as well as the DFG Cluster of Excellence:
Munich-Center for Advanced Photonics. The authors
would like to thank Magnus Bergh, Gerard Kleywegt, In-
ger Andersson, Erik G. Marklund, Richard Neutze, Mar-
tin Svenda, Rosmarie Friemann, Jochen Hub and Karin
Valegard for their valuable input.
[1] L. F. DiMauro, J. Arthur, N. Berrah, J. Bozek, J. N.
Galayda, and J. Hastings. Progress report on the LCLS
XFEL at SLAC. J. Phys.: Conf. Ser., 88:012058, 2007.
[2] Coline Nave and Mark A. Hill. Will reduced radiation
damage occur with very small crystals? J. Sync. Rad.,
12:299 -- 303, 2005.
[3] John A. Cowan and Colin Nave. The optimum conditions
to collect x-ray data from very small samples. J. Sync.
Rad., 15:458 -- 462, 2008.
[4] M. Bergh, N. Tımneanu, and D. van der Spoel. A model
for the dynamics of a water cluster in a X-ray FEL beam.
Phys. Rev. E, 70:051904, 2004.
[5] C. Caleman, C. Ortiz, E. Marklund, F. Bultmark,
M. Gabrysch, F. G. Parak, J. Hajdu, M. Klintenberg,
and N. Tımneanu. Radiation damage in biological mate-
rial: electronic properties and electron impact ionization
in urea. Europhys. Lett., 85:18005, and erratum 88:29901,
2009.
[6] Z. Jurek, G. Faigel, and M. Tegze. Dynamics in a cluster
under the influence of intense femtosecond hard X-ray
pulses. Eur. Phys. J. D, 29:217 -- 229, 2004.
[7] Stefan P. Hau-Riege, Richard A. London, and Abraham
Szoke. Dynamics of biological molecules irradiated by
short x-ray pulses. Phys. Rev. E, 69:051906, 2004.
[8] Beata Ziaja, Richard A. London, and Janos Hajdu. Uni-
fied model of secondary electron cascades in diamond. J.
Appl. Phys., 97:064905, 2005.
[9] Carlos Ortiz and Carl Caleman.
Secondary electron
cascade dynamics in KI and CsI. J. Phys. Chem. C.,
111:17442 -- 17447, 2007.
[10] Richard A. Kirian, Xiaoyu Wang, Uwe Weierstall,
Kevin E. Schmidt, John C. H. Spence, Mark Hunter,
Petra Fromme, Thomas White, Henry N. Chapman, and
James Holton. Femtosecond protein nanocrystallogra-
phydata analysis methods. Opt. Express, 18:5713 -- 5723,
2010.
[11] R. Neutze, R. Wouts, D. van der Spoel, E. Weckert, and
J. Hajdu. Potential for biomolecular imaging with fem-
tosecond X-ray pulses. Nature, 406:752 -- 757, 2000.
[12] J. Miao, K. O. Hodgson, and D. Sayre. An approach
to three-dimensional structures of biomolecules by using
single-molecule diffraction images. Proc. Natl. Acad. Sci.
U.S.A., 98:6641 -- 6645, 2001.
[13] Stefan P. Hau-Riege. Effect of the coherence properties
of self-amplified-spontaneous-emission x-ray free electron
lasers on single-particle diffractive imaging. Opt. Express,
16(4):2840 -- 2844, 2008.
[14] IV Tomov, P Chen, and PM Rentzepis. Pulse broad-
ening in femtosecond x-ray diffraction. J. Appl. Phys.,
83(10):5546 -- 5548, MAY 15 1998.
[15] David van der Spoel, Erik Lindahl, Berk Hess, Gerrit
Groenhof, Alan E. Mark, and Herman J. C. Berendsen.
GROMACS: Fast, Flexible and Free. J. Comp. Chem.,
26:1701 -- 1718, 2005.
[16] N. Tımneanu, C. Caleman, J. Hajdu, and D. van der
Spoel. Auger electron cascades in water and ice. Chem.
Phys., 299:277 -- 283, 2004.
[17] M. Gabrysch, E. Marklund, J. Hajdu, D. J. Twitchen,
J. Rudati, A. M. Lindenberg, C. Caleman, R. W. Fal-
cone, T. Tschentscher, K. Moffat, P. H. Bucksbaum,
J. Als-Nielsen, A. J. Nelson, D. P. Siddons, P. J. Emma,
P. Krejcik, H. Schlarb, J. Arthur, S. Brennan, J. Hast-
ings, and J. Isberg. Formation of secondary electron
cascades in single-crystalline plasma-deposited diamond
upon exposure to femtosecond x-ray pulses. J. Appl.
Phys., 103:064909, 2008.
[18] David van der Spoel, Filipe R. N. C. Maia, and Carl
Caleman. Structural studies of melting on the picosec-
ond time scale. Phys. Chem. Chem. Phys., 10:6344 -- 6349,
2008.
[19] ZH Kalman. On the derivation of integrated reflected
energy formulae. Acta Cryst. A, 35:634 -- 641, JUL 1979.
[20] A. J. C. Wilson. The probability distribution of X-ray
intensities. Act. Cryst., 2:318 -- 321, 1949.
[21] Don T. Cromer and Joseph B. Mann. X-ray scatter-
ing factors computed from numerical Hartree-Fock wave
functions. Acta Cryst. A, 24:321 -- 326, 1968.
[22] R Fourme, A Ducruix, M Rieskautt, and B Capelle. The
perfection of protein crystals probed by direct record-
ing of bragg reflection profiles with a quasi-planar x-ray
wave. J. Sync. Rad., 2(Part 3):136 -- 142, MAY 1 1995.
|
1312.3450 | 1 | 1312 | 2013-12-12T11:27:35 | Scarcity may promote cooperation in populations of simple agents | [
"physics.bio-ph",
"physics.soc-ph",
"q-bio.PE"
] | In the study of the evolution of cooperation, resource limitations are usually assumed just to provide a finite population size. Recently, however, it has been pointed out that resource limitation may also generate dynamical payoffs able to modify the original structure of the games. Here we study analytically a phase transition from a homogeneous population of defectors when resources are abundant to the survival of unconditional cooperators when resources reduce below a threshold. To this end, we introduce a model of simple agents, with no memory or ability of recognition, interacting in well-mixed populations. The result might shed light on the role played by resource constraints on the origin of multicellularity. | physics.bio-ph | physics |
Scarcity may promote cooperation in populations of simple agents
1Departament de Física, Universitat Autònoma de Barcelona, Campus UAB, E-08193 Bellaterra, Spain.∗
(Dated: January 20, 2020)
R. J. Requejo1 and J. Camacho1
In the study of the evolution of cooperation, resource limitations are usually assumed just to
provide a finite population size. Recently, however, it has been pointed out that resource limitation
may also generate dynamical payoffs able to modify the original structure of the games. Here we
study analytically a phase transition from a homogeneous population of defectors when resources
are abundant, to the survival of unconditional cooperators when resources reduce below a threshold.
To this end, we introduce a model of simple agents, with no memory or the ability of recognition,
interacting in well-mixed populations. The result might shed light on the role played by resource
constraints on the origin of multicellularity.
PACS numbers: 02.50.-r,87.10.-e,87.23.-n,89.75.Fb
Cooperation is common in nature in all levels of biolog-
ical organization [1], and it is considered to have played a
key role in the evolutionary appearance of higher selective
units, such as eukaryotic cells or multicellular life, from
simpler components [2]. However, its widespread abun-
dance is intriguing because cooperators are vulnerable to
exploitation by defectors [3], as detected early on [4, 5].
Since then, several mechanisms have been found allowing
cooperative behaviors to survive such as network struc-
ture, group selection, direct and indirect reciprocity or
tag-based donation [6, 7]. Behavioral mechanisms -- the
latter three are examples of such mechanisms -- require
players to have some ability to avoid the exploitation
from defectors, such as memory or the capacity to recog-
nise the co-player [6, 7]. As a result, simple agents with-
out these abilities, such as unconditional cooperators, are
not expected to survive in well-mixed populations.
Aside from a few examples [8 -- 11], the role played by
the limitation of resources in most evolutionary game the-
oretical studies on the origin and persistence of coopera-
tion has been just to impose a constant population size
[6, 7, 12 -- 17]. Recently, however, we have put forward a
new viewpoint where the interacting players are set into
a nonequilibrium context [18 -- 20]. The environment is
considered explicitly by introducing a resource flux into
the system that drives it away from equilibrium. This
standpoint leads to unexpected outcomes, such as that
resource limitation allows for stable coexistence between
unconditional cooperators and defectors, and even dom-
inance of cooperation, in well-mixed populations playing
an a priori Prisoner's Dilemma (PD) game. This hap-
pens due to a self-organizing process involving the envi-
ronment which generates dynamical payoffs transforming
the original PD structure into a neutral game.
One of the main results of the analysis performed in
ref.[19, 20] is that a well-mixed population of uncon-
ditional cooperators extinguishes for infinite resources
(where the system plays a PD game) but may survive for
some parameter values when resources are finite (where
the game is not a PD anymore). This suggests the possi-
bility of a transition from a population of only defectors
when resources are abundant to a population containing
cooperators for more stringent environments. The exis-
tence of this transition should have great interest, since
it would provide a resource-based mechanism preventing
the spread of defectors and thus may shed light on the
conditions under which cooperators could survive during
the evolutionary process. Indeed, the survival of cooper-
ative strains has been recently observed experimentally
in yeast (S.Cerevisiae) [21 -- 23] and bacterial (E.Coli)
[24, 25] cultures, and has also been found in a model
for the survival of aerobic cells inside anaerobic cultures
[26, 27]. The models depicted in refs.[19, 20], however, do
not yield such a transition: in these models we considered
that the population was ruled by a resource limiting re-
production, and that deaths occurred at a constant rate,
so that the limiting resource influx determined the pop-
ulation size; as it was thoroughly discussed, a reduction
in the resource flux just decreased the size of the popu-
lation in the same proportion, but it did not modify its
composition.
Our aim here is to devise a scenario where the selection
pressure drived by resource limitation combined with the
nonlinearities induced by this resource limitation in the
interactions among players may lead to a transition of
the type discussed above. In this scenario, we assume a
limiting resource that constraints reproduction in a pop-
ulation of constant size due, for instance to space con-
straints. The plaussibility of the latter assumption is dis-
cussed at the end of the paper. The model developed here
is a stylized one inspired in the model of ref. [20], which
consists of an evolving population of self-replicating indi-
viduals that receive resources from the environment and
exchange resources during interactions. In order to avoid
the effect of spatial structure and focus on the effect of re-
source constraints, we consider a well-mixed population.
No memory, learning abilities or any other sensory in-
puts are assumed. Each individual i is represented by its
amount of resources, Ei, which in this simplified model
is either 0 or 1, and its strategy, namely cooperate (C) or
defect (D). Its amount of resources may be interpreted as
the amount that belongs to it independently of how (it
may be in its surroundings, for instance). Each defector
attacks at a rate α per unit time to individuals chosen
at random and steals its internal resources. To do so,
the defector must have internal resources greater than
0 (i.e. Ei = 1), otherwise it does not attack. In every
interaction, the defector loses its unit of resources with
probability q, which can thus be seen as the average cost
paid by a defector in an interaction. If the interaction
partner has no resources, no reward is obtained. Coop-
erators do nothing, they just eventually suffer from de-
fectors' attacks. We assume that behaviors are inherited
without mutation and represent physiologic or morpho-
logic characteristics intrinsic to individuals which cannot
be modified by choice.
Each individual receives from the environment γ units
of resources per unit time independently of its strat-
egy, thus not modifying the interaction payoff structure.
When an individual with internal resources Ei = 1 re-
ceives an extra unit of resources it splits into two identical
copies, each one with Ei = 1. Along with reproduction,
we assume that players die with a probability f , indepen-
dently of its strategy, in such a way that the number of
individuals in the population remains constant. There-
fore, resource allocation, reproduction and death rules
are equal for both cooperators and defectors, being the
strategy the only difference.
Let us note that, in this model, an increase in the envi-
ronmental resource supply is represented by an increase
in γ, the amount of per capita resources obtained by in-
dividuals. This contrasts with the model in refs. [19, 20],
where an increase of resources leads to a proportional in-
crease in the population size while keeping the same per
capita value.
We consider simultaneous interactions and large pop-
ulations so that we can make a continuum approach. We
denote by c0 and c1 the fraction of cooperators with in-
ternal resources 0 and 1, and d1 and d0 = 1 − c0 − c1 − d1
the fraction of defectors with internal resources 1 and 0,
respectively. The equations governing the evolution of
cooperators are the following
dc0
dt
dc1
dt
= αc1d1 − γc0 − f c0
= −αc1d1 + γ(c0 + c1) − f c1
(1)
(2)
The αc1d1 term shows the fraction of cooperators C1 that
lose their internal resource unit after the attack of defec-
tors (the latter pertaining to the population d1); these
individuals move from population c1 to c0. The term in
γc0 quantifies the fraction of individuals C0 that change
to population c1 after getting a unit of resources from
the environment. In addition, individuals in population
c1 that receive resources from the environment replicate,
thus increasing the c1 population. The terms f ci describe
2
the fraction of individuals dying in each population per
unit time.
To describe the evolution of defectors is enough to
write the equation for population d1 because d0 is just
the remaining fraction of the whole population. The dy-
namic equation for d1 is
dd1
dt
= −αqd1 + αc1d1 + γ(d0 + d1) − f d1.
(3)
The terms related to deaths and resource allocation from
the environment are analogous as for cooperators. The
interaction term is as follows. On the one hand, with
probability q individuals D1 lose its resource unit when
interacting with individuals C0, D0 and D1; this leads
to a decrease in the population of d1 in an amount
αqd1(c0 + d0 + d1). On the other hand, when interacting
with individuals C1, individuals D1 sequester their re-
source unit; therefore, either the population of D1 does
not change, with probability q, or it increases due to re-
production at a rate αc1d1(1 − q).
To complete the equations of the model, we need an
expression for the death rate f . In order to have a con-
stant population size, the frequency of deaths must equal
the frequency of reproductions. This leads to
f = γ(c1 + d1) + α(1 − q)(c1 + d1)d1.
(4)
The first term denotes reproductions due to resource al-
location and the second one to reproduction of D1 in-
dividuals when attacking individuals with Ei = 1 and
not paying the cost. Eqs. (1) -- (4) are the equations of
our model. They can be further simplified by noticing
that one can divide all the equations by parameter γ
and absorb it into the time parameter; therefore, there
are just two dimensionless parameters in the model, q
and β = α/γ. A large β value indicates either large de-
fector attack rates or small resource influxes from the
environment; conversely, large resource influxes or small
attacking rates yield small β values. The dimensionless
equations are the same Eqs (1) -- (4) replacing α by β, and
γ by 1.
The numerical resolution of the model shows that the
system is attracted to a globally stable fixed point in-
dependent of initial conditions. Depending on the pa-
rameter values, the final fate is either a population of
defectors (an expected solution) or, interestingly, a sta-
ble mixture of cooperators and defectors. Remarkably
enough, for fixed q, small β values, i.e. large resource in-
fluxes, provide a population of just defectors, but when
β exceeds a critical value βc a mixed state appears, thus
providing a smooth phase transition from defective states
to mixed states as resources become scarce (see Fig. 1a).
The existence of stable mixed states in the model may
be explained in terms of the overexploitation mechanism
discussed in ref. [19]: an excess of defectors may reduce
cooperators' resource contents and, as a result, the av-
erage reward obtained by defectors; eventually, rewards
(a)
l
s
a
u
d
v
d
n
i
i
i
f
o
n
o
i
t
c
a
r
f
(b)
r
'
E
,
t
i
f
e
n
e
b
'
s
r
o
t
c
e
f
e
d
c
n
o
i
t
c
e
f
e
d
l
l
u
F
Coexistence
0
10
20
30
40
50
attack probability to resource income ratio,
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
0.0
0.2
0.4
0.6
defectors' cost, q
0.8
1.0
FIG. 1: (a) Phase transition for q=0.5. The fraction of coop-
erators c0, c1 and defectors d1 (see text) above the threshold
are denoted with solid, dashed and dotted lines, respectively.
Below the critical value βc = 7.58 cooperators die out. (b)
Defectors' benefit versus costs in coexistence states. It equals
the frequency rate function f (q) (see text). The dashed line
E ′
r = q is a guide to the eye.
decrease below costs and cooperators recover. Interest-
ingly, we can obtain simple analytical expressions for the
composition of the mixed state as a function of parameter
β above the threshold:
ci = ai(1 −
βc
β
),
d1 =
a2
β
,
(5)
with ai and βc functions of parameter q.
Remarkably, the dynamics in coexistence states selfor-
ganizes defectors' rewards to be (almost) equal to costs
thus turning the payoff matrix to neutral. According to
the model, the payoff matrix for an average interaction
is
C
0
D
−E ′
C
D E ′
(cid:18)
r − Ec −Ec (cid:19)
r
(6)
with E ′
r the average reward obtained when a defector at-
tacks a cooperator, and Ec = q the average cost paid
when a defector attacks. Then, the average reward re-
ceived by defectors when interacting with cooperators is
3
r = c1/(c0 + c1). Eq. (5) shows that E ′
E ′
r = a1/(a0 + a1)
and then it is a function dependent only on q, and not
on β. Fig. 1b displays the reward E ′
r as obtained nu-
merically versus the cost q showing that E ′
r ≃ q. They
are not exactly equal because, as explained in Eq. (4) of
ref. [19], they may differ when death frequencies f are
not small compared with resource intake. In this model,
f cannot be arbitrarily chosen because of the constant
population condition. Indeed, Eqs. 1 -- 2 readily show that
f = c1/(c0 +c1) = E ′
r and then Fig. 1b also displays f (q).
One observes that f is generally of order 1 (this is the
cause of the small deviations found in Fig. 1b). At small
q, however, f is also small and E ′
r and q match perfectly.
One can further study the transition by drawing a
phase diagram β − q with the regions where each be-
havior dominates. It is possible to obtain an analytical
expression for the critical curve βc(q) by performing a
stability analysis. To do so, let us recall that for a fixed
point to be stable in three dimensions the trace and de-
terminant of the Jacobian matrix must be negative. Our
model system (1) -- (4) has at least two fixed points, cor-
responding to pure populations of cooperators and de-
fectors: (A) c1 = 1 (the remaining variables equal to 0),
and (B) c0 = c1 = 0 and d1 6= 0 obeying, according to
Eq. 3,
d1(f + βq) = 1.
(7)
As we know, it may also have a mixed fixed point, given
by Eq. 5, but it need not be considered for our present
purpose. Linear stability of fixed point A leads to the
jacobian matrix
−2 0
β
1 −1 −1 − 2β + βq
−1 −1 β(1 − q) − 1
,
(8)
with determinant D = 4β(1 − q) > 0. The positive sign
shows that at least one of its three eigenvalues is positive.
Then point A is always unstable for q < 1 and coopera-
tors never occupy the whole population. Fixed point B
provides the following jacobian matrix
−1 − f∗
βd∗
1
1
1 − βd∗
1 − f ∗
0
0
,
df
dd1
−1
−1 − d∗
1
+ βd∗
1 −βq − f∗ − d∗
1
df
dc1
(9)
where d∗
1 and f∗ are the values of these quantities in fixed
point B. To be compact, let us call J33 = −(βq + f∗ +
d∗
1)) < 0. The trace is T = J33 − 2f∗ −
βd∗
1 < 0, and the determinant can be written as
1(1 + 2β(1 − q)d∗
D = J33(−1 + βd∗
1f∗ + f 2
∗ ).
(10)
e
c
r
u
o
s
e
r
o
t
y
t
i
l
i
b
a
b
o
r
p
k
c
a
t
t
a
,
o
i
t
a
r
e
m
o
c
n
i
50
40
30
20
10
0
0.0
Coexistence
Defection
0.2
0.4
0.6
0.8
1.0
defectors' cost, q
FIG. 2: Phase diagram. The solid line indicates the numerical
solution, the dashed line the analytical approximation βc =
q−3. Cooperation is favoured at large β, i.e. small resource
fluxes or large attacking rates (see text).
Then, for point B to be stable the term inside parenthe-
sis has to be positive. Although this is not a sufficient
condition to prove that point B is stable, the numerical
resolution of Eqs. (1) -- (4) shows that this is the case; this
is the region where defectors are dominant. When the
parenthesis in Eq. (10) is negative point B becomes un-
stable, wich means that a small fraction of cooperators
will grow and survive (notice that point B is the only
fixed point with only defectors). Then, since point A is
also unstable, in this situation there must exist a third
(mixed) fixed point in the dynamics. Eq. (5) supplies
the solution for this mixed fixed point and numerical so-
lutions show it is a stable attractor, the one describing
the stationary coexistence of cooperators and defectors
found at large β values. In order to obtain the cuve βc(q)
separating the regions of dominance of defectors from the
mixture of cooperators and defectors we should find d∗
from Eqs. (4) and (7) and solve the equation
1
− 1 + βcd∗
1f∗ + f 2
∗ = 0.
(11)
The exact analytical solution of this transition curve is
very cumbersome, so that we try two alternative routes.
One is to obtain a numerical solution (see Fig.2), the
other one is to find an approximate analytical solution.
In this sense, let us note that, if β 2q3 ≫ 1 Eqs. (4) and
1 ≃ (βq)−1, because βq ≫ f∗ ≃ (βq2)−1.
(7) show that d∗
In this limit, the instability condition (11) just gives
βc = q−3,
(12)
which provides an excellent approximation not only for
c q3 ≃ q−3 ≫ 1 (say q <
β 2
∼ 0.5 ) but over the whole range
0 ≤ q ≤ 1 as shown when compared with the exact nu-
merical solution (Fig. 2).
Fig. 2 shows that cooperation is favored at large costs
q and large β, whereas defectors dominate in the opposite
4
limit. The origin of the dependence on the average cost
q is rather direct: the larger the cost, the less favorable
for defectors to reproduce. The dependence on param-
eter β is, however, counterintuitive since (at first sight)
one would expect that large attack rates (large β) should
benefit defectors. The explanation is not easy due to the
nonlinearities involved in the model. One might think
that the origin of the observed behavior relies on the ex-
ploitation mechanism that explains the existence itself
of coexistence states, and accordingly reason that large
attacking rates would cause a great damage on coopera-
tors, which would reduce rewards over costs, ultimately
harming defectors. However, this is not what happens,
since we have seen above that the average reward E ′
r is
a function of q only, and then it does not change when
increasing β at fixed q.
One explanation of why large β favor cooperators is
that it leads to a small fraction of defectors in the active
state (D1), thus reducing the damage on cooperators. In
effect, if resources are abundant individuals receive them
frequently and there will be large populations of D1 indi-
viduals; if resources are scarce, only a few individuals will
be in state Ei = 1. The same occurs if attacking rates are
large. Since attacks are indiscriminate, defectors are also
victim of the attacks, which decrease the number of D1
individuals; conversely for small attacking rates. This ex-
planation is consistent with the behavior of d1 displayed
in Fig. 1a. Indeed d1 decreases yet from β = 0, i.e. below
the transition, as it can be seen from our approximate so-
1 ≃ (βq)−1. Below some critical population value
lution d∗
depending on q (around q2) the reduced population of de-
fectors in the active state is not capable of extinguishing
cooperators. In this point is worthwhile to point out that
parasites continuously receive resources from the environ-
ment and interact, and then, they change from active to
inactive states continuosly. In the stationary state, the
fraction of defectors in the population is d0 + d1. These
defectors spend a fraction of time d0/(d0 + d1) in inactive
states and d1/(d0 + d1) in active states.
In summary, we have developed a simple model de-
scribing a phase transition from defective parasitic pop-
ulations (which dissipate some amount of resources in
order to gain a higher reproduction rate) when resources
are abundant to the survival of cooperators when re-
sources are scarce. This is the result of a self-organizing
process involving subtle nonlinearities in the interactions
induced by resource constraints. In contrast to previous
models, where the same limiting resource ruled repro-
duction and population size [19, 20] and do not display
this transition, the model studied here assumes that the
factor limiting reproduction is different from the one lim-
iting the population size, so that populazion size remains
approximately constant. This may be accomplished in
chemostat or retentostat experiments [28 -- 31], which may
allow for laboratory testing of the predictions of the
model. Indeed, recent experiments with S.Cerevisiae [21]
and E.Coli [24] at low concentration of glucose agree with
the results presented here of the survival of cooperative
traits instead of their expected extinction in unstructured
populations.
In natural environments, the constant population as-
sumption may apply in situations where space constraints
the size of the population more restrictively than re-
source scarcity. Of course, a complete description of spa-
tially distributed populations goes beyond the mean-field
model presented here, and should consider that interac-
tions occur only among neighbors. Note however, that
space alone is well-known to favor cooperation because
it permits the formation of clusters of cooperators. Since
we have seen here that resource limitation alone already
allows for the survival of cooperation, the combined ef-
fects of both space and resource limitations are expected
to enhance the conditions under which cooperators can
prevent their extinction. Other extensions of the model,
beside the inclusion of space, may be the introduction
of continuous behaviors (and not just two, namely coop-
erate or defect), what could shed light on the observed
phenotypic radiation of behaviors in E.Coli [25], which
represents an exception to the competitive exclusion prin-
ciple [33].
In a broader scope, the results presented here might
have played a role in the route towards the emergence
of multicellularity by cooperative aggregation triggered
by resource constraints. It has been argued [22, 26, 27]
that such transition happened whenever cooperative indi-
viduals formed clusters, which subsequently evolved nu-
trient exchange between the components of the cluster,
and later evolved a joint replication mechanism, stage
at which a higher-order organism can be considered to
exist. However, such studies do not explain why coop-
erative bacteria survived in a first stage before forming
clusters. The mechanism presented here provides some
insights for the maintainance of such cooperative indivi-
dals before clusters of cooperators could form, being a
first step towards the formation of multicellular organ-
isms.
This work has been supported by the Spanish gov-
ernment (FIS2009-13370-C02-01) and the Generalitat
de Catalunya (2009SGR0164). R.J.R. acknowledges
the financial support of the Universitat Autònoma de
Barcelona and the Spanish government (FPU grant).
∗ Electronic address: [email protected]
[1] P. Hammerstein, ed., Genetic and Cultural Evolution of
Cooperation (Dahlem Workshop Report 90) (Cambridge,
5
Massachusetts: MIT Press., 2003).
[2] J. Maynard Smith and E. Szathmary, The Major Tran-
sitions in Evolution (Freeman, Oxford, 1995).
[3] C. Darwin, On the Origin of Species by Means of Natural
Selection (John Murray, London, 1859), 1st ed.
[4] W. D. Hamilton, J. Theor. Biol. 7, 1 (1964).
[5] W. D. Hamilton, J. Theor. Biol. 7, 17 (1964).
[6] R. L. Riolo, M. D. Cohen, and R. Axelrod, Nature 414,
441 (2001).
[7] M. A. Nowak, Science 314, 1560 (2006).
[8] J. Y. Wakano, J. Theor. Biol. 247, 616 (2007).
[9] U. Dobramysl and U. C. Tauber, Phys. Rev. Lett. 101,
258102 (2008).
[10] C. Hauert, M. Holmes, and M. Doebeli, Proc. Roy. Soc.
London B 273, 2565 (2006).
[11] A. Melbinger, J. Cremer, and E. Frey, Phys. Rev. Lett.
105, 178101 (2010).
[12] R. L. Trivers, Q. Rev. Biol. 46, 35 (1971).
[13] R. Axelrod and W. D. Hamilton, Science 211, 1390
(1981).
[14] M. A. Nowak and K. Sigmund, Nature 393, 573 (1998).
[15] A. Traulsen and H. Schuster, Phys. Rev. E 68 (2003).
[16] J. Gomez-Gardenes, M. Campillo, L. M. Floria, and
Y. Moreno, Phys. Rev. Lett. 98 (2007).
[17] C. P. Roca, J. A. Cuesta, and A. Sanchez, Phys. Rev. E
80 (2009).
[18] R. J. Requejo and J. Camacho, J. Theor. Biol. 272, 35
(2011).
[19] R. J. Requejo and J. Camacho, Phys. Rev. Lett. 108,
038701 (2012).
[20] R. J. Requejo and J. Camacho, Phys. Rev. E 85, 066112
(2012).
[21] M.L.A. Jansen, J.A. Diderich, M. Mashego, A. Hassane,
J.H. de Winde, P. Daran-Lapugadel and J.T. Pronkl, Mi-
crobiol. 151, 1657-1669 (2005)
[22] J. H. Koschwanez, K. R. Foster, and A. W. Murray, Plos
Biology 9, e1001122 (2011).
[23] J.W. Wenger, J. Piotrowski, S. Nagarajan, K. Chiotti,
G. Sherlock, and F. Rosenzweib, Plos Genet. 7, e1002202
(2011).
[24] L. Notely-McRobb and T. Ferenci, Environm. Microbiol.,
45-52 (1999)
[25] R. Maharjan, S. Sheeto, L. Notely-McRobb and T. Fer-
enci, Science., 514-517 (2006)
[26] T. Pfeiffer, S. Schuster, and S. Bonhoeffer, Science 292,
504 (2001), ISSN 0036-8075.
[27] T. Pfeiffer and S. Bonhoeffer, Proc. Nat. Ac. Sci. USA
100, 1095 (2003), ISSN 0027-8424.
[28] A. Novik and L. Szilard, Science 112, 715-716 (1950)
[29] J. Monod, Annales De L Institut Pasteur 79, 390-410
(1950)
[30] H.W. Jannasch, Limnol. Oceanogr. 19, 716-720 (1974)
[31] A. Groisman, C. Lobo, H.J. Cho, J.K. Campbell,
Y.S. Dufour, A.M. Stevens and A. Levchenko Nature
Methods 2, 685-689 (2005)
[32] J. Alonso, A. Fernandez, and H. Fort, J. Stat. Mech.,
P06013 (2006)
[33] G. Hardin, Science, 1292 (1960)
|
0911.2722 | 2 | 0911 | 2011-05-20T21:49:58 | A Stochastic Compartmental Model for Fast Axonal Transport | [
"physics.bio-ph",
"math.PR",
"q-bio.NC"
] | In this paper we develop a probabilistic micro-scale compartmental model and use it to study macro-scale properties of axonal transport, the process by which intracellular cargo is moved in the axons of neurons. By directly modeling the smallest scale interactions, we can use recent microscopic experimental observations to infer all the parameters of the model. Then, using techniques from probability theory, we compute asymptotic limits of the stochastic behavior of individual motor-cargo complexes, while also characterizing both equilibrium and non-equilibrium ensemble behavior. We use these results in order to investigate three important biological questions: (1) How homogeneous are axons at stochastic equilibrium? (2) How quickly can axons return to stochastic equilibrium after large local perturbations? (3) How is our understanding of delivery time to a depleted target region changed by taking the whole cell point-of-view? | physics.bio-ph | physics |
A Stochastic Compartmental Model for Fast
Axonal Transport
Scott A. McKinley∗
Lea Popovic†
Michael C. Reed‡
October 29, 2018
Abstract
In this paper we develop a probabilistic micro-scale model and use
it to study macro-scale properties of axonal transport, the processes by
which materials are moved in the axons of neurons. By directly mod-
eling the smallest scale interactions, we can use recent microscopic ex-
perimental observations to infer all the parameters of the model. Then
using techniques from queueing theory, we can predict macroscopic
behavior in order to investigate three important biological questions:
(1) How homogeneous are axons at stochastic equilibrium? (2) How
quickly can axons return to stochastic equilibrium after large local
perturbations? (3) How inhomogeneous does deposition and turnover
make the axon?
1
Introduction
In all cells, one finds that proteins, membrane-bound organelles, and other
structures (e.g. chromosomes) are transported from place to place at speeds
much higher than diffusion. Though these transport processes are funda-
mental to cell function, many of the underlying mechanisms, organizational
principles, and regulatory features remain unknown. Axonal transport is
one of the best studied systems because the transport is basically one-
dimensional since axons are long and narrow. There are two speeds of axonal
transport. Fast transport goes at speeds of roughly 0.2 to 0.5 meters/day
∗Department of Mathematics, Duke University, Box 90320, Durham, NC 27701
([email protected])
†Department of Mathematics and Statistics, Concordia University, Montreal QC H3G
1M8
‡Department of Mathematics, Duke University, Box 90320, Durham, NC 27701
1
2
[20][24], while slow transport goes at approximately 1 millimeter/day, the
rate of axon growth and regeneration [5][20]. The biology and principles of
slow transport are not yet clear [5], but the basic mechanisms of fast ax-
onal transport were discovered in the 1980s [2][3][22][34]. The model in this
paper refers to fast axonal transport, which we will henceforth call axonal
transport.
The axonal transport apparatus consists of vesicles which form reversible
chemical bonds with motor proteins that bind reversibly to microtubules
which run parallel to the long dimension of the axon [1]. When the vesicle-
motor protein complex is assembled on the microtubule, the complex steps
stochastically with step size approximately 8 nanometers for kinesin and
dynein and 10 nanometers for myosin [6][10][16][31]. The vesicles enter from
the cell body on microtubules and then detach and reattach to the transport
mechanism at random times.
In this paper we propose a spatial Markov-chain compartmental model
based on these dynamics. We will assume independence of the interactions,
and exponential wait times between events. While we address the validity of
these assumptions in the Discussion section, we consider this a useful "first-
order" approximation that permits study of the dynamics from both the
perspective of individual vesicles as well as that of the full spatial system.
Such a model unifies some earlier modeling efforts and can accommodate
both qualitative and quantitative experimental data observed on multiple
scales.
In much experimental work in the 1970s and 1980s, radio-labeled amino
acids were put into the cell bodies continuously or for a few hours. The amino
acids were incorporated into proteins that were packaged into vesicles and
put on the transport system so that at later times radioactivity could be
seen moving progressively down the axons. In the continuous infusion case,
one would see a wave of radioactivity with a sharp but slowly spreading
wavefront propagating at constant velocity down the axon. In the case of
infusion for a few hours one would see at long times a slowly spreading pulse
of radioactivity that looked normally distributed. It was to understand this
behavior that Reed and Blum constructed PDE models for axonal transport
[3][25][26]. These models did not have traveling wave solutions, but the data
certainly looked like approximate traveling waves. In [27] it was shown by
a perturbation theory argument that, in the asymptotic limit where the
unbinding and binding rates k1 and k2 get large, the solution approaches
a slowly spreading traveling wave or a normal pulse. Recently, in a series
of papers, Friedman and co-workers have introduced new PDE models and
proved these results rigorously[12][13][14][15].
3
Probabilistic models for axonal transport were introduced and used for
simulations already in the 1980s [30][32]. However, rigorous work began with
Lawler [21] in 1995 and was continued by Brooks [4] who used a continuous
time stochastic model to show that the distribution of an individual particle
is a spreading Gaussian at large times. Brooks also proved tail estimates for
the central limit theorem and used them to estimate the error from normal.
1.1 Summary of Results
In this paper we revise the existing probabilistic models in order to ob-
serve randomness in the system as a whole rather than exclusively from the
particle's point of view. Our goal is to exploit experimental observations
at multiple scales in order to make predictions about the perceived homo-
geneity of material along the length of the axon. To this end, we propose
in Section 2 a continuous-time Markov chain queueing model for the axonal
transport system and estimate the order of magnitude of the various param-
eters. The precise values of the parameters will vary for different molecular
motors, for various types of cargo and for various animal species.
In Section 3 we take the individual vesicle point-of-view to recover pre-
vious results [4] [26] and show that this model does unify and extend the
existing probabilistic and PDE models. Using standard results from re-
newal theory we calculate the mean velocity and the near-Gaussian wave-
front spreading of the law of the vesicle's location. Since we assume indepen-
dence of the particle interactions, the law of an individual is equivalent to the
distribution of an ensemble of particles released at the same time. Therefore
the PDE governing a spatial-continuum limit of the law of a single particle
is the same as the PDE studied in [26] (see Section 3.3).
Subsequently in Section 4 we adopt the full spatial system perspective to
quantify stochasticity along the length of the axon. We begin by calculating
in Proposition 4.1 the stationary distribution of a flow-through system that
has sustained input from the nucleus, while particles are removed upon
reaching the distal end. The stationary distribution has a product Poisson
structure which allows for seamless transition between spatial scales. Via
the coefficient of variation we give a precise characterization in Section 4.2 of
the experimenter's qualitative perception of homogeneity at the millimeter
scale.
The final two sections deal with model predictions.
In Section 5 we
study the non-equilibrium dynamics of recovery. Due to the product struc-
ture of the law of the transient dynamics, behavior is determined by the 2N -
dimensional ODE governing the means. From this we estimate the timescale
4
of return to equilibrium as a function of the lengthscale of interest. In Sec-
tion 6 we consider the effect of deposition of material in the cell membrane.
We find that there should be an exponential loss of material along the length
of the cell, and conclude that a new biological mechanism must be discov-
ered to account for the homogeneity that is a hallmark of axonal transport
experimental observations.
2 The model and its parameters
Let L be the length of the axon, divided evenly into N lateral sections each
of length δ, equal to the step size of the motor protein. Within each section,
we disregard any further spatial geometry and take the particles to be in
one of two states:
• an on-transport state with mean lateral velocity v, or
• an off-transport state with lateral velocity 0.
We use a 2N -dimensional continuous time Markov chain to model the
particle dynamics:
• Qi(t) is the number of particles in the on-transport state in section i,
• Pi(t) is the number of particles in the off-transport state in section i.
Each Qi and Pi has the natural numbers N = {0, 1, 2, . . .} as its state space.
In the simplest model, we consider the following transitions and rates,
• Lateral transport:
(Qi, Qi+1) → (Qi − 1, Qi+1 + 1) at rate rQi(t), where r = v/δ;
• Switch from on-transport to off-transport:
(Qi, Pi) → (Qi − 1, Pi + 1) at rate k1Qi(t);
• Switch from off-transport to on-transport:
(Pi, Qi) → (Pi − 1, Qi + 1) at rate k2Pi(t);
• Production of new particles:
(Q1) → (Q1 + 1) at rate q0r;
• Removal of particles at distal end:
(QN ) → (QN − 1) at rate rQN (t).
Q1 = 3
P1 = 3
∗
•
∗
∗
∗
•
x = L
3
Q2 = 1
P2 = 2
∗
•
•
N
u
c
l
e
u
s
x = 0
5
Q3 = 3
P3 = 2
•
∗
∗
•
•
x = 2L
3
x = L
Q1
✲q0r
Q2
✲r
Q3
✲r
✲r
❑
k2
k1
❯
P1
❑
k2
k1
❯
P2
❑
k2
k1
❯
P3
Figure 1: Double Chain
The lateral transport rate, r, is inversely proportional to the length scale so
that the mean number of particles per unit length is invariant with respect
to rescaling δ. The rate of production, q0r, ensures that this mean number
per unit scales with q0. A graph of the model is depicted in Figure 1.
In order to ensure the Markov property, we use exponential random
variables for the waiting times between transition events. Specifically, we
mean that after a given event we assign a new independent random variable
to each of the 3N + 2 possible next events, exponentially distributed with
the appropriate rate parameter. The system of values updates according to
the transition associated with the minimum of these waiting times. Then we
create a new set of exponential random variables and the process proceeds
as before.
6
This is an exactly solvable model. The advantage of computing explicit
formulas for quantities that can be observed in experiments is that the ex-
perimental data can then be used to determine the parameter values in the
model. For the characterization of the approximate wavefront speed and
spreading in Section 3 and the homogeneity calculations in Section 4 we
need order-of-magnitude estimates for the parameters. Actual parameter
values will certainly differ depending on the particular neural tissue and
the particular particles being transported. However, we can get order-of-
magnitude estimates from existing data.
First we recall that fast transport has been observed to travel at speeds of
0.2 to 0.5 m per day. We can assume that the average velocity of particles
while physically bound to microtubules is roughly 1 m per day, or 10−6
m/s. We have already stated the assumption that the length scale of the
individual steps satisfies δ ∼ 10−8m. This implies that the rate parameter
should be r = v/δ = 102s−1.
We now turn our attention to the on-off rates rates k1 and k2. These can
be determined from experimentally observed run lengths on the transport
system. Indeed, Dixit et al. [9] show that a typical run along microtubules
for dinein and kinesin is on the order of 10−6m. We can compare this with
the theoretical run length of the model to determine off-rate k1. Within the
model, at each step on the transport mechanism the particle has a binary
decision to jump laterally along the transport with probability r/(k1 + r),
or to jump off with probability k1/(r + k1). The number of jumps along the
transport system before jumping off is therefore geometrically distributed
. It follows that average
on the set {0, 1, . . .} with success probability
number of steps in the run is r
and therefore the average run length is
k1
r
k1 × 10−8 m. Setting this equal to the average experimental run length of
10−6 from [9], we see that r
k1 ∼ 102, implying that k1 ∼ 1s−1. As we will see
in the computation of the stationary distribution in Section 4.1, the ratio of
the expected number of particles on the track to those off the track is k1
.
k2
Dixit [9] found that approximately 75% of the particles were motile so this
ratio is approximately equal to 3. Since k1 ∼ 1s−1 we see that k2 ∼ 1
3 .
It remains to estimate q0. We will see in Proposition 4.1 that the mean
number of particles per unit length is (1 + k1
)q0 = 4q0. Of course, axons
k2
have a large variety of diameters and larger axons will have more vesicles per
unit length so one expects a range of values for q0. However, examination
of a large number of electron micrographs of axonal cross-sections (see for
example [17], Fig. 3; [18]; [23]), which are typically 100 nm thick enables
one to estimate the number of vesicles per 100 nm segment. This number
r+k1
r
7
is typically in the range of 10 to 100 which implies that there are 1 to 10
vesicles per "box" in our model. Therefore q0 is in the range 0.25 to 2.5, for
various axons.
We remark that we are ignoring some aspects of the physics and the
biology of axonal transport. We are not including diffusion of the vesicles
off the track. We are treating the microtubule track as though it were
a single continuous entity from one end of the axon to the other, when
it fact it consists of numerous, separated, microtubule fragments. And,
we are ignoring retrograde transport and the details of the motor proteins.
Nevertheless, this simple model will enable us to investigate the homogeneity
questions that are the main goal of this paper.
3 Dynamics from the Particle Perspective
In this section we calculate properties of the stochastic dynamics by using
standard theorems from queuing theory. In the δ → 0 limit, the law of the
location of a single particle converges to the Green's function of a linear
partial differential equation. This enables us to obtain, as a special case,
the asymptotic behavior of the PDE models for axonal transport in various
asymptotic limits.
3.1 The active transport mode
We first consider the simple case where the particle starts at X0 = 0 and
stays exclusively in active transport mode. Let Xt ∈ {0, δ, 2δ, . . . , L} be the
lateral position of a particle at time t and let nt be the number of jumps
made by the particle as of time t. Observe, Xt = δnt.
Proposition 3.1. Let k1 = k2 = 0, and r = v/δ > 0, then Xt ∼ P ois(vt).
In particular, the mean velocity of the particle is given by 1
E[Xt] = v. For
t
any given t ≥ 0, in the limit as δ → 0 the position of the particle satisfies
Xt
t
a.s.−→ v and
1
√δ(cid:16)Xt − vt(cid:17)t≥0
where B is a standard Brownian motion.
d=⇒√v(cid:0)Bt(cid:1)t≥0
Proof. Since Xt = δnt where (cid:0)nt(cid:1)t≥0 is a Poisson process with rate r = v/δ
the set out results follow from the law of large numbers (LLN) and the
central limit theorem (CLT) for a Poisson process.
8
3.2 The on/off dynamics
We now consider a particle which undergoes transitions from on-transport
to off-transport state and back. Denote again by Xt ∈ {0, δ, 2δ, . . . , L} the
lateral position of a particle at time t and let nt be the number of lateral
transition jumps made by the particle as of time t. Observe that the particle
will spend only a fraction of its time in active transport and hence the lateral
speed of the particle should be slower than before.
Proposition 3.2. Let k1, k2 > 0, and let r = v/δ > 0, then the mean
velocity of the particle satisfies 1
t
k1+k2
v.
k2
E[Xt] −→t→∞
Proof. Consider the time τ for a particle to make one step of lateral trans-
port. Before doing so a particle performs m switches from on-transport
k1+r )
to off-transport and back, where m is distributed as a Geometric(
variable. The amounts of time a particle takes for each switch τ Q→P
from
on-transport to off-transport are iid variables with Exponential(k1+r) distri-
bution. Likewise, the amounts of time a particle takes for each switch τ P→Q
from off-transport back to on-transport are iid variables with Exponential(k2)
distribution, and are independent of the other switching times. Then
r
i
i
m
τ =
i
Xi=1(cid:0)τ Q→P
+ τ P→Q
i
m+1
(cid:1) + τ Q→Q
with E[τ ] = (k1 + k2)/(k2r). Let nt be the number of times a particle
makes a step of lateral transport until time t. By the Renewal Theorem
1
.
t
E[τ ] , and the result follows from Xt = δnt and δ k2r
= k2v
k1+k2
k1+k2
1
E[nt] →t→∞
Proposition 3.3. Let k1, k2 > 0, and r = v/δ > 0. In the limit as δ → 0
the position of a particle satisfies
Xt
t
a.s
−→t→∞
k2
k1 + k2
v, and √t(cid:16) Xt
t −
k2
k1 + k2
=⇒t→∞s 2k1k2
v(cid:17) d
(k1 + k2)3 vB1
where B1 ∼ Normal(0, 1).
Proof. The number of lateral transport steps (nt)t≥0 is a renewal chain with
E[τ ] = (k1 + k2)/k2r and Var[τ ] = ((k1 + k2)2 + 2k1r)/r2k2
2. By the LLN
and CLT for Renewal chains
nt
t
a.s−→t→∞
1
E[τ ]
, and √t(cid:16) nt
t −
1
E[τ ](cid:17) d=⇒t→∞s Var[τ ]
(E[τ ])3 B1
and now the result follows from Xt = δnt, δ 1
E[τ ] = δ k2r
k1+k2
9
= k2
k1+k2
v, and
δ2 Var[τ ]
(E[τ ])3 =
δk2v
k1 + k2
+
2k1k2v2
(k1 + k2)3 ≈
2k1k2
(k1 + k2)3 v2 for δ ≈ 0.
The fact that an individual particle starting at x = 0 will have the distri-
bution given by Proposition 3.3 at long times was also obtained by Brooks
[4] who proved tail estimates on the CLT and found that the remainder is
O( 1√t
). In our approach, one can obtain the same error estimate using the
asymptotic analysis of renewal chains, see, for example, [7].
3.3 Connection to Partial Differential Equations Models.
In order to demonstrate the connection between our model and the PDEs
seen in [27][26][14], we make a formal calculation regarding convergence of
the law of the location of a single particle in the full system.
Let δN = 1
N and let XN (t) denote the lateral position of a particle in a
system with N boxes. Define
qN (x, t) := P{XN (t) = x, and the particle is on track}
pN (x, t) := P{XN (t) = x, and the particle is off track.}
We suppress N in the notation. By definition of the Poisson process, for
small values of h we have
q(x, t + h) = q(x, t)(1 − (k1 + r)h) + q(x − δ, t)rh + p(x, t)k2h + o(h)
p(x, t + h) = p(x, t)(1 − k2h) + q(x, t)k1h + o(h).
which we may reorganize as
1
(q(x, t + h) − q(x, t)) = r[q(x − δ, t) − q(x, t)] − k1q(x, t) + p(x, t)k2 +
h
1
(p(x, t + h) − p(x, t)) = −k2p(x, t) + k1q(x, t) +
h
Formally, we take the Taylor expansion of q in space: q(x − δ, t) = q(x, t) −
δ∂xq(x, t) + o(δ2). Taking limits as h → 0 and δ → 0 yields the system of
PDEs
o(h)
h
o(h)
.
h
∂tq(x, t) + v∂xq(x, t) = −k1q(x, t) + k2p(x, t)
∂tp(x, t) = k1q(x, t) − k2p(x, t).
(1)
(2)
10
When k1 = 0 = k2, the limiting PDE is simple linear transport: (∂t +
v∂x)q(x, t) = 0. The initial condition q(x, 0) = δ0(x) corresponds to the
density of a single particle at the origin at t = 0. The time evolution via
simple linear transport is translation of the delta function, while the time
evolution via the equations (1) and (2) will have a spreading profile. This is
seen in Propositions 3.1 and 3.3 since the variance of the fluctuations goes
to 0 in Proposition 3.1 but not in Proposition 3.3 as δ → 0.
In the experiments described in the introduction one sees "approximate"
traveling waves of radioactivity in the axons in the sense that there is a slowly
spreading wave front moving at constant velocity away from the cell body.
The equations (3) and (4) are linear and do not have solutions that are
bounded traveling waves. However, it was shown by a perturbation theory
argument in [27][26] that as ε → 0 the solutions of
ε(∂t + v∂x)qε(x, t) = −k1qε(x, t) + k2pε(x, t),
ε∂tpε(x, t) = k1qε(x, t) − k2pε(x, t).
qε(0, t) = q0
(3)
(4)
are to leading order
qε(x, t) = H(
x − at
ε1/2 , t),
pε(x, t) = cH(
x − at
ε1/2 , t),
where H satisfies the heat equation
∂sH(y, s) =
κ2
2
∂yyH(y, s),
H(y, 0) = χ(−∞,0),
and
a =
k1v
k1 + k2
,
κ2 =
2k1k2v2
(k1 + k2)3 .
(5)
This asymptotic form is valid for small ε, that is for large k1 and k2. How-
ever, if we set q(x, t) = qε( x
ε ), then q and p satisfy
(3) and (4), so the solutions of (3) and (4) behave like approximate travel-
ing waves for large t and large x whether or not k1 and k2 are large. The
results suggested by these perturbation theory arguments have been proven
rigorously by Friedman and coworkers [12][13][14][15].
ε ) and p(x, t) = pε( x
ε , t
ε , t
In the case that radiolabeled particles enter the axon only for a short
time, 0 ≤ t ≤ T , the same perturbation analysis shows that for large times
the solution of (3) and (4), where q(0, t) = 0 for t > T is to leading order a
spreading Gaussian with mean and variance as given in Proposition 3.3. This
11
corresponds to what is seen experimentally and arises asymptotically in the
solutions of (3) and (4) by convolving the asymptotic solution corresponding
to the initial condition q(x, 0) = δ0(x) with the indicator function χ[0,T ](x).
4 Dynamics from the spatial system perspective
4.1 The spatial system in equilibrium
We are now ready to characterize the steady state dynamics induced by
continually adding particles from the nucleus and removing them when they
reach the distal end of the cell.
Proposition 4.1. If the incoming rate of particles at Q0 is q0r, then the
system has the following stationary distribution
Qi ∼ P ois(q0),
k2 (cid:19)
Pi ∼ P ois(cid:18) k1q0
where the {Qi} and {Pi} are mutually independent.
Proof. Given the value of the input rate Q0 = q∗ the generator of the process
(Q1, P1) in the first section is
Aq∗f (q, p) = [f (q + 1, p) − f (q, p)]rq∗ + [f (q − 1, p) − f (q, p)]rq
+ [f (q − 1, p + 1) − f (q, p)]k1q + [f (q + 1, p − 1) − f (q, p)]k2p
If we first use f (q, p) = Q1(t) then use f (q, p) = P1(t) and take expectations
we get a system of ODE's governing the change in E[Q1], E[P1] over time
dE[Q1](t)
dt
dE[P1](t)
dt
= rE[Q0] − rE[Q1(t)] − k1E[Q1(t)] + k2E[P1(t)]
= k1E[Q1(t)] − k2E[P1(t)]
indicating that in equilibrium E[Q1] = E[Q0], E[P1] = E[Q1] k1
. For pur-
k2
poses that will soon be clear we first assume that Q0 has a constant Pois(λ∗)
distribution over time rather than just being equal to λ∗. Let π(q∗, q, p) =
πλ∗(q∗) ⊗ πλQ(q) ⊗ πλP (p) where πλ are Pois(λ) distributions with rates
λ∗, λQ = λ∗, and λP = λ∗
respectively. To show that π(q∗, q, p) is a
stationary distribution for the process (Q1, P1) we need to check that
k1
k2
∞
Xq∗=0
∞
Xq=0
∞
Xp=0
Aq∗f (q, p)π(q∗, q, p) = 0
12
q∗
q∗
∞
∞
∞
∞
∞
λp
P
λp
P
p!
λq
Q
q!
λq
Q
q!
Xp=0
λ∗
q∗!
for any choice of function f ∈ D(Aq∗).
Aq∗f (q, p)e−(λ∗+λQ+λP ) λ∗
Xq∗=0
Xq=0
q∗!
= e−(λ∗+λQ+λP ) ∞
Xq∗=0
Xp=0
p! (cid:0)[f (q + 1, p) − f (q, p)]rq∗ + [f (q − 1, p) − f (q, p)]rq
+ [f (q − 1, p + 1) − f (q, p)]k1q + [f (q + 1, p − 1) − f (q, p)]k2p(cid:1)(cid:17)
In the inner two sums, for a fixed value of q∗, the factor multiplying f (q, p)
for any q, p ∈ N2 comes only from terms involving {q − 1, q, q + 1} and
{p − 1, p, p + 1} and equals e−(λ∗+λQ+λP ) times
λp
P
p!
λq+1
Q
(q + 1)!
Xq=0
rq∗ −
λq
Q
q!
λp
P
p!
λp
P
p!
(cid:16)
rq
λq−1
Q
(q − 1)!
r(q + 1) −
λq
Q
q!
λq
Q
q!
λp
P
p!
λp
P
p!
k1q
k2p
k1(q + 1) −
k2(p + 1) −
λQ
q + 1
p
λP
k1(q + 1)
λp
P
p!
rq∗ +
λp−1
P
(p − 1)!
λp+1
P
(p + 1)!
λq
Q
q!
λq+1
Q
(q + 1)!
λq−1
Q
(q − 1)!
rq∗ − rq∗ +
λP
p + 1
q
λQ
λQ
q + 1
+
+
λq
Q
q!
=
λp
P
p! (cid:16) q
λQ
r(q + 1) − rq +
k2(p + 1) − k2p(cid:17)
=
λp
P
λq
Q
q!
− k1q +
rq∗ − rq∗ + λ∗r − rq(cid:17)
p! (cid:16) q
since λQ = λ∗ and λP
Summing over q∗ the factor multiplying f (q, p) becomes
= k1
k2
λ∗
λQ
.
e−(λQ+λP ) λq
Q
q!
λp
P
p!
= e−(λQ+λP ) λq
Q
q!
∞
q∗
rq∗ − rq∗ + λ∗r − rq(cid:17)
λ∗
e−λ∗ λ∗
q∗! (cid:16) q
Xq∗=0
p! (cid:16)qr − rλ∗ + λ∗r − rq(cid:17) = 0
λp
P
13
We've just shown that if the distribution of input particles Q0 in equi-
librium is Pois(λ∗) then the stationary distribution of Q1, P1 is independent
of the distribution of Q0 and is πλQ ⊗ πλP where λQ = E[Q0], and λP =
E[Q0]k1/k2. It is easy to see in the above calculation that if the input Q0 was
indeed constant q0 then the stationary distribution of (P1, Q1) would again
be πλQ ⊗ πλP with rates λQ = q0 and λP = q0k1/k2 respectively. Now, since
the stationary distribution of Q1 is the input into (Q2, P2) process, the more
general calculation shows that in equilibrium the stationary distribution of
(Q2, P2) is independent of the stationary distribution for Q1 and is πλQ⊗πλP
It
with rates λQ = E[Q1] = q0 and λP = E[Q1]k1/k2 = q0k1/k2 again.
follows by induction that the stationary distributions for {(Qi, Pi)} are in-
dependent and identically distributed as πλQ ⊗ πλP , λQ = q0, λP = q0k1/k2.
This is also an example of a clustering process satisfying the detailed balance
conditions with linear rates discussed in Sec. 8.2 of [19].
4.2 Homogeneity of the axons at equilibrium
Recall that δ = 10nm, roughly the step size of motor proteins, and that
axons can be up to one meter in length. Thus we are interested in phenomena
on all the length scales 10ν δ, where ν = 1, 2, . . . 8. Let ∆ = 10ν δ; we want to
determine how similar different segments of the axon of size ∆ are. Let Q∆
and P∆ denote the numbers of on-track and off-track particles in a segment
of length ∆.
In equilibrium, Q∆ and P∆ are both sums of 10ν independent Poisson
random variables with parameters λQ = q0 and λP = k1
q0, respectively.
k2
Therefore the distributions of Q∆ and P∆ are Poisson with parameters
10ν λQ and 10ν λP , respectively. The mean and the variance of the num-
ber of particles in the segment of length ∆ is 10ν (λQ + λP ). To see how
homogeneous different slices of length ∆ are, we consider the coefficient of
variation, c∆, which is the standard deviation divided by the mean.
c∆ =
1
=
p(λQ + λP )10ν
p(1 + k1/k2)q0∆108
1
.
As indicated in Section 2, q0 is in the range 0.25 to 2.5 in different axons.
For illustrative purposes here, we will assume q0 = 1. Since k1/k2 = 3,
we see that the scale-dependent coefficient of variation c∆ = 1/(2√108∆).
Therefore at the ten nanometer scale the coefficient of variation is simply
1/2. At the micron scale c∆ = 1/20 and at the millimeter scale c∆ = 0.5 ×
10−5/2. The cutoff between "high variance" and "low variance" distributions
14
is usually considered to be when the coefficient of variation is near 1, so by
this standard the axon is extremely homogeneous in its length at large scales.
5 Approaching Equilibrium
We have seen above that the axon is very homogeneous at stochastic equilib-
rium on a space scale down to micrometers. One of the beautiful properties
of the system of transport with reversible binding is that if it is locally out
of equilibrium, the dynamics will automatically take it back to equilibrium.
This is of fundamental importance to the biological function of the system
because it means that the axon will automatically "repair" itself without
central control of the repair process. How good is this mechanism? If a
segment of the axon is far away from equilibrium, how long does it take
to get back close to equilibrium? We study this question first for a single
location and then use the estimates derived to scale the results to segments
of any length.
Proposition 5.1. Let (Q, P )(0) = (0, 0) and let the constants a > 0 and
p ∈ (0, 1) satisfy the relationship a2pλ∞ > 1 where λ∞ is the equilibrium
vector of (Q, P )(t). Then there exists a t∗ such that t ≥ t∗,
In fact, the choice
is sufficient, where
α :=
P{(Q, P )(t) − λ∞ ≥ aλ∞} ≤ p.
t∗ = α−1 ln(cid:16) ppλ∞
appλ∞ − 1(cid:17)
1
2(cid:16)k1 + k2 + r −p(k1 + k2 + r)2 − 4k2r(cid:17)
Proof. We begin by noting that for any given β ∈ (0, 1), we may choose
t∗ > 0 such that for all t ≥ t∗, the vector of means λ(t) := E[(Q, P )(t)]
satisfies λ(t) − λ∞ ≤ aβλ∞. Then
P{(Q, P )(t) − λ∞ ≥ aλ∞} ≤ P{(Q, P )(t) − λ(t) + λ(t) − λ∞ ≥ aλ∞}
≤ P{(Q, P )(t) − λ(t) ≥ a(1 − β)λ∞}
Applying Chebyshev's Inequality, and observing that the variance of a Pois-
son random variable is equal to its mean, we conclude that
P{(Q, P )(t) − λ∞ ≥ aλ∞} ≤
Var[(Q, P )(t)]
a2(1 − β)2λ∞2 =
λ(t)
a2(1 − β)2λ∞2
15
Since the initial condition for both P and Q are less than their respec-
tive equilibrium values, each are monotonically increasing functions and the
above reduces to
P{(Q, P )(t) − λ∞ ≥ aλ∞} ≤
1
a2(1 − β)2λ∞
for all t > t∗. In order to satisfy the requirement that the right hand side
must be less than p, we solve for β and find
1
appλ∞
β = 1 −
provided that appλ∞ > 1.
It remains to study the convergence of the mean and the appropriate
choice of t∗. The dynamics of the mean vector λ(t) are given by the ODE
(6)
λ(t) = −A1λ(t) + q0re1
where e1 is the unit vector (1, 0) and
A1 =(cid:18) k1 + r −k2
k2 (cid:19) .
−k1
The solution to (6) is
λ(t) = λ∞ + e−A1t(λ(0) − λ∞)
where λ∞ = q0rA−1
1 e1 = q0(1, rk1
k2
). This yields the estimate
where α is the smaller of the eigenvalues of A1, namely
λ(t) − λ∞ ≤(cid:12)(cid:12)e−A1tλ∞(cid:12)(cid:12) ≤ e−αtλ∞
(k1 + k2 + r −p(k1 + k2 + r)2 − 4k2r).
Noting that α > 0, t∗ may be chosen so that e−αt∗ ≤ aβ, i.e.
α =
1
2
t∗ = α−1 ln(1/(aβ))
16
We now suppose that the whole axon is in statistical equilibrium except
for a segment of length δ × 10ν m in which we will assume that there are no
particles either on or off the track. We want to know how long will it take
for this segment to get back close to equilibrium. Of course, Proposition
5.1 covered the case ν = 0. We are interested in ν = 1, . . . , 8. We imagine
that the axon is broken up into 108−ν segments of length δ × 10ν m. In this
rescaled system, the unbinding and binding rates per particle, k1 and k2
remain the same, as well as the mean on-transport velocity v. In order to
retain this mean velocity, the rate of lateral stepping must be decreased to
r = r × 10−ν .
The ODE for the mean vector of the rescaled system is given by
d
dt
λ(t) = − A1 λ(t) + q0re1
(7)
with
A1 =(cid:18) k1 + r −k2
k2 (cid:19) .
−k1
We note that the last term in (7) contains an r rather than an r. This is
because the input rate is unchanged while the exit rate is diminished.
The resulting equilibrium value is therefore rescaled as well,
λ∞ = q0r A−1
1 e1 = q0
r
r(cid:18) 1
k1/k2(cid:19).
Both components of this vector are of order 10ν , as expected.
We will use the parameters discussed in Section 2: k1 = 1, k2 = 1
3 , v =
10−6m/s, r = 102s−1. This implies r = 102−ν s−1. As in our analysis in
Proposition 5.1, the time to equilibrium, t∗, is proportional to α−1 where α
satisfies:
1
α =
=
2(cid:16)k1 + k2 + r −p(k1 + k2 + r)2 − 4k2r(cid:17)
k1 + k2 + r +p(k1 + k2 + r)2 − 4k2r
2k2r
.
We are most interested in the cases ν ≥ 3 (the segment has length ≥ 10 mi-
crons). Then k2r is small compared to k1. Ignoring constants and restricting
our attention to the leading order terms gives
α ∼
k2r
k1 + k2 + r ∼ 102−ν s−1
17
Thus,
t∗ ∼ 10ν−2s.
Thus, for a 10 micron segment (ν = 3) the time to equilibrium is about
10 seconds and for a 1 millimeter segment (ν = 5) the time to equilibrium is
1000 seconds or 15 minutes. Of course, the time to get close to equilibrium
depends on the parameter p that represents what we mean by "close."
6 Transport with Deposition
One major feature of the biology we have ignored is that some particles in
the off-transport state are actually deposed permanently in the membrane
or are used for some other purpose. Such a phenomenon is easy to add to
the model proposed above, but we must address a new qualitative feature
of the results. Namely there is an exponential loss of material as we move
toward the distal end.
Q1
✲q0r
Q2
✲r
Q3
✲r
✲r
❑
k2
k1
❯
P1
❑
k2
k1
❯
P2
❑
k2
k1
❯
P3
k3
❄
k3
❄
k3
❄
D1
❄1/τ
D2
❄1/τ
D3
❄1/τ
Figure 2: Fast Axonal Transport with Deposition
18
In addition to deposition, we recognize that particles are only useful in
the membrane for a random amount of time with a given half-life, after which
they decompose and get transported back to the cell body. Neglecting this
retrograde transport, we augment the previous model as follows: at each δ-
length lateral section let Di(t) be the number of particles deposed in section
i, with N = {0, 1, 2, . . .} as the state space for Di. In addition to the previous
dynamics we also define the following transition rates
• (Pi, Di) → (Pi − 1, Di + 1) (deposition) at rate k3Pi(t)
• Di → Di − 1 (decay) at rate 1/τ where τ > 0
The constant τ in the decay rate is the average lifetime of a deposed particle.
6.1 Dynamics from the particle perspective
As before, we can use the renewal theorem to determine the average speed
of particles that are not deposed, as well as the spreading of the wavefront.
E[Xt] −→t→∞
Proposition 6.1. Let k1, k2, k3 > 0, τ < ∞ and let r = v/δ > 0, then the
mean velocity of the particle that is not deposed into the membrane by time
t satisfies 1
t
Moreover, in the limit as δ → 0 the position of a particle satisfies
Xt
t
k2 + k3
=⇒t→∞s 2k1(k2 + k3)
v(cid:17) d
(k1 + k2 + k3)3 vB1
k1 + k2 + k3
v, √t(cid:16) Xt
t −
k1 + k2 + k3
a.s
−→t→∞
k2+k3
k1+k2+k3
v.
k2 + k3
where B1 ∼ Normal(0, 1).
Remark 1. Notice the mean velocity of particles is higher than when there
is no deposition
k2 + k3
(k1 + k2 + k3)
v ≥
k2
(k1 + k2)
v, with equality iff k3 = 0
A simple explanation for this is that particles that are still in lateral trans-
port had returned to on-transport back from off-transport before managing
to be deposed into the membrane, meaning that their time off-transport was
shorter than it would have been if they did not have to beat the exponential
clock for deposition.
Proof. When there is deposition then by time t each particle has either
already deposed, or is still performing lateral transport. A particle that is
deposed into the membrane has velocity zero from the time of its deposition
on. A particle that can still be laterally transported must have avoided
deposition until time t.
19
If we proceed along the lines of the proof without deposition, then the
only difference now is that the amount of time a particle takes for each switch
from off-transport back to on-transport are i.i.d. exponential variables with
parameter k2+k3 rather than k2. Hence the amount of time τ for a particle to
make a step from one section to another has E[τ ] = (k1+k2+k3)/((k2+k3)r),
and by the Renewal Theorem the number of steps a particle that can still
be laterally transported satisfies 1
t
(k1+k2+k3) r.
(k2+k3)
E[nt] →t→∞
Also Var[τ ] = ((k1 + k2 + k3)(k1 + k2 + k3 + 2k1r))/r2(k2 + k3)2, so the
result now follows by the Renewal LLN and CLT for nt and the fact that
δ
1
E[τ ]
= δ
(k2 + k3)r
k1 + k2 + k3
=
k2 + k3
k1 + k2 + k3
v,
and
δ2 Var[τ ]
(E[τ ])3 =
δ(k2 + k3)v
k1 + k2 + k3
+
2k1(k2 + k3)v2
(k1 + k2 + k3)3 ≈
2k1(k2 + k3)
(k1 + k2 + k3)3 v2 for δ ≈ 0
6.2 Dynamics from the spatial system perspective
Once again we compute the stationary distribution of the flow-through sys-
tem, this time with deposition.
Proposition 6.2. Suppose the incoming rate of particles at Q0 is q0r. Then
the stationary distribution of the system is
q0ρi(cid:19)
Qi ∼ P ois(q0 ρi), Pi ∼ P ois(cid:18) k1
where ρ = r(k2 + k3)/((r + k1)(k2 + k3) − k1k2) and {Qi}, {Pi}, and {Di}
are mutually independent.
q0ρi(cid:19) , Di ∼ P ois(cid:18) k1k3τ
k2 + k3
k2 + k3
Proof. Since (Qi, Pi) do not depend on Di the proof for the stationary dis-
tribution of (Qi, Pi) follows along the same lines as in the case when there
is no deposition. We first check the stationary distribution for the process
(Q1, P1) in the first section. For given value of Q0 = q∗ the generator of the
20
process (Q1, P1) is given by
Aq∗f (Q1, P1) = [f (Q1 + 1, P1) − f (Q1, P1)]rq∗ + [f (Q1 − 1, P1) − f (Q1, P1)]rQ1
+ [f (Q1 − 1, P1 + 1) − f (Q1, P1)]k1Q1 + [f (Q1 + 1, P1 − 1) − f (Q1, P1)]k2P1
+ [f (Q1, P1 − 1) − f (Q1, P1)]k3P1
The system of ODE's governing the change in E[Q1], E[P1] is now
∂tE[Q1]
∂t
∂tE[P1]
∂t
= E[Q0]r − E[Q1]r − E[Q1]k1 + E[P1]k2
= E[Q1]k1 − E[P1]k2 − E[P1]k3
Hence, in equilibrium
and
E[Q1] = E[Q0]
k2 + k3
(r + k1)(k2 + k3) − k1k2
= q0ρ
E[P1] = E[Q1]
k1
k2 + k3
=
k1q0
k2 + k3
ρ.
If we let λQ = E[Q1], λP = E[P1], and define the distribution π(Q0, Q1, P1) =
πq0δ(Q0) ⊗ πλQ(Q1) ⊗ πλP (P1), then P∞q∗,q,p=0 Aq∗f (q, p)π(q∗, q, p) = 0 can
be verified exactly as before. Since now the mean of Q1 differs from the
mean of Q0 by a multiplicative factor ρ < 1, and this is the mean of the
input into Q2, it follows by induction that the sequence of rates for {Qi, Pi}
Poisson random variables decays geometrically in i by a factor ρi. This
immediately implies that in equilibrium the magnitude of the variances also
decays geometrically along the line of progression of lateral transport.
To find the stationary distribution for Di note that if the number of
particles at Pi = p∗ the generator for Di is
Ap∗f (d) = [f (d + 1) − f (d)]k3p∗ + [f (d − 1) − f (d)](1/τ )d
It is easily checked that if Pi ∼ πλPi
for Di which is independent of the distribution for Pi with λDi = k3τ λPi.
is the stationary distribution
then πλDi
Corollary 6.3. When in equilibrium the mean and the variance of the num-
ber of particles along the axon decay exponentially, with a loss ratio
E[Qi+1]
E[Qi]
=
Var[Qi+1]
Var[Qi] ≈ exp(cid:18)−
δk1
k2 + k3
k3
v (cid:19)
(8)
and the same loss ratio holds for the off-transport and the deposed particles.
21
Remark 2. Notice that the rate of loss is depends on the ratio of reaction
rates k1/k2, which we earlier estimated to be approximately 3, as well as
the ratio k3/v, and that if k1, k2 ≫ k3 then the loss rate is e−
Proof. Recalling ρ = r(k2 + k3)/(r(k2 + k3) + k1k3) and r = v/δ, we rewrite
3k3
v .
ρ =(cid:16)1 + δ
k1k3
v(k2 + k3)(cid:17)−1
Observing that for any small δ, the approximation (1 + δx)−1 = 1 − δx +
(δx)2 − . . . ≈ e−δx, we have Equation (8).
The same ratio holds for the other two states as well
E[Pi]/E[P0] = Var[Pi]/Var[P0] = E[Di]/E[D0] = Var[Di]/Var[D0] = ρi
6.3 Homogeneity despite deposition
It remains to estimate the rates k3 and τ from experimental data and then
make a formal calculation to address the volume of material loss due to
deposition and turnover as a function of length scale. To be precise, suppose
we want to know the percentage loss over a length ∆ = 10ν δ. Then from
the Corollary 6.3 the fraction of material loss is approximately
E[Q10ν ]
E[Q1] ≈
10ν
Yi=1
exp(cid:16) − δ
k1
k2 + k3
k3
v (cid:17) = exp(cid:16) − 10ν−8
k1
k2 + k3
k3
v (cid:17)
(9)
So, for example, at the meter scale, ν = 8, and so the loss ratio is exp(−k1k3/v(k2+
k3)).
We now consider the parameters of a specific example, sodium pumps
being deposed in the cell membrane. We assume that the half life of a single
pump is approximately a week, which means τ = 6 × 105s. It remains to
estimate k3. According to [36], there are roughly 1000 sodium pumps per
micron2 of membrane surface area. If we take the diameter of the neuron to
be 5 microns, then the membrane surface area for one 10 nm slice is about
0.15 µm2. So there are roughly 150 sodium pumps per slice.
It is important to note that each vesicle carries roughly 100 sodium
pumps and so the stationary distribution for the quantity Di is on average
1/100th of the number of sodium channels which are deposed into the mem-
brane at the location i. Taking this into account and recalling our earlier
parameter estimations, k1 = 1, k2 = 1/3, v = 10−6 and q0 = 1 we estimate
the average number of sodium pumps in the membrane of the first 10 nm
slice to satisfy 150 = 100 q0k1k3τ /(k2 + k3) which implies k3 = 1 × 10−6.
From the above we see that the loss ratio at the meter scale is
22
exp(cid:16) −
k1
k2 + k3
k3
v (cid:17) = e−3 ≈ 0.04
(10)
which is to say that over the length of an entire meter long axon, we expect
the concentration of sodium pumps in the membrane near the distal end to
be roughly a 1/25 of that found near the nucleus. It is striking to note that
over a length 0.1 m, however, the loss ratio is e−.3 = .74. This implies that
material loss due to deposition and turnover becomes an issue exactly over
the range of lengths of typical axons: 0.1 m to 1 m. At the lower end of
this range, the axon is very homogeneous. At the higher end, one predicts
significant inhomogeniety.
7 Discussion
In this paper, we used a spatial Markov chain model to unify several pre-
vious approaches to modeling fast axonal transport. After inferring the
orders of magnitude of the parameters of the model, we have used standard
techniques from renewal theory and linear Markov chain theory to give a
precise description of the most striking qualitative feature of axonal trans-
port: homogeneity along the length of the axon. Furthermore we predict the
timescale of recovery to this homogeneous state after removing the vesicles
from a segment as a function of the length scale of the segment.
In creating the model, we have become aware of two features of the
biology which have not been adequately studied: 1) Deposition and decay
of proteins; and 2) characterization of the process by which unbound vesicles
reattach to the transport mechanism.
Regarding deposition and decay, we mentioned in Section 6 that the
model unequivocally predicts a loss of material down the length of the axon.
Furthermore, it seems that certain biological features that we have chosen
to ignore -- such as retrograde transport and a location-dependent unbinding
rate [9] -- would only serve to increase the non-homogeneity. In the case of
sodium pumps, it is not immediately clear if there is a contradiction between
deposition loss and homogeneity, but we do see a sensitive dependence on
the length of the axon. Furthermore, it stands to reason that for other types
of cargo the loss may significant. We believe that this raises an interesting
biological issue, and mention a few possible mechanisms for overcoming the
23
material loss: first, it is possible that proteins can be synthesized in axons
[33]; second, there may be some signal that prevents material from deposing
in a region that is already "full;" and third, it may be that for some materials
"deposition" is a reversible process.
Regarding reattachment, we note that we have modeled all event wait
times as exponential random variables, but this is certainly a simplification.
As an example, the stepping process of kinesin is a well-studied though still
not completely understood phenomenon. Much work has focused on as-
sessing the dependence of the mean rate of translocation on both the load
and the local concentration of ATP [35] [29].
Implicit in this analysis is
the assumption of exponential wait times with state dependent rate param-
eters. However, when fitting to data and matching dispersion information
the authors in [11] found it necessary to generalize the wait time distribu-
tion. This was followed by more detailed models for which it was shown that
load carrying could in fact regularize the stepping times of kinesin motors
[28] [8].
Generalizing the wait times does not significantly affect the particle per-
spective results of Sections 3.2 and 6.1. This is because renewal theory,
which depends on independence of the wait times, is robust with respect to
such changes. However, the spatial system results may be affected in that
the full process is no longer Markov. While one may still expect a some-
thing like a stationary distribution for the flow-through system (where new
particles enter from the nucleus and particles are removed from the system
at the distal end), the analysis of the approach to equilibrium may change.
In the absence of direct observation of the phenomenon, we refrain from this
more detailed analysis.
In light of this known need for generalized wait times in the stepping
process, it seems likely that detailed observation of the rebinding process
will call for new models as well. Recall that when a vesicle unbinds from
a microtubule it is unclear whether it typically rebinds to the same micro-
tubule or if it explores the region significantly via diffusion before finding
a different microtubule to bind to. In the latter case, a more appropriate
model for rebinding time would be to solve some kind of first passage time
problem and use that distribution for the rebinding wait.
Acknowledgments
The authors are grateful to Professors H. Frederik Nijhout and Vann Ben-
nett of Duke University and Professor Anthony Brown of The Ohio State
24
University for helpful discussions. This work was supported by NSF Grant
DMS-061670.
References
[1] Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2008)
The Molecular Biology of the Cell, Garland Science, Taylor & Francis
Group, New York.
[2] Allen RD, Weiss DG, Hayden JH, Brown DT, Fijiwaki H, Simpson
M (1985). Gliding movement and bidirectional transport along single
native microtubules from squid axoplasm: evidence for an active role
for microtubules in cytoplasmic transport. J. Cell Biol 100, 1736-1752.
[3] Blum JJ, Reed MC (1985) A Model for Fast Axonal Transport. Cell
Motility 5, 507-527.
[4] Brooks, E. A. (1999) Probabilistic methods for a linear reaction-
hyperbolic system with constant coefficients. Ann. Appl. Prob., 9(3),
719-731.
[5] Brown A (2000) Slow axonal transport: stop and go traffic in the axon.
Nat Rev Cell Mol Biol 1, 153-156.
[6] Carter NJ, Cross RA (2005) Mechanics of the kinesin step. Nature 435,
308-312.
[7] Cox DR (1962) Renewal Theory. Methuen & C0. Ltd., London, Section
4.5.
[8] DeVille RL and Vanden-Eijnden E (2008) Regular Gaits and Optimal
Velocities for Motor Proteins Biophysics Journal 95 6: 2681-2691.
[9] Dixit R, Ross JL, Goldman YE, Holzbaur ELF (2008) Differential
regulation of dynein and kinesin otor proteins by tau. Science 319,
1086-1089.
[10] Finer JT, Simmons RM, Spudich JA (1994) Single myosin molecule
mechanics: piconewton forces and nanometre steps. Nature 368, 113-
119.
[11] Fisher ME and Kolomeisky (2001) Simple mechanochemistry describes
the dynamics of kinesin molecules. PNAS 98 14: 7748-7753.
25
[12] Friedman A, Craciun G (2005) A model of intracellular transport of
particles in an axon. SIAM J Appl Math 38, 741-758.
[13] Friedman A, Craciun G (2006) Approximate traveling waves in linear
reaction-hyperbolic equations. SIAM J Appl Math 38, 741-758.
[14] Friedman A, Hu B (2007) Uniform convergence for approximate trav-
eling waves in reaction-hyperbolic systems. Indiana Math J 56, 2133-
2158.
[15] Friedman A, Hu B (2007) Uniform convergence for approximate trav-
eling waves in reaction-diffusion-hyperbolic systems. Arch Rat Mech
Anal 186, 251-274.
[16] Gennerich A, Carter AP, Reck-Peterson, Vale RD (2007) Force-induced
bidirectional stepping of cytoplasmic dynein. Cell 131, 952-965.
[17] Gross, G.W, Weiss, D.G. (1982) "Theorestical considerations on rapid
transport in low viscosity axonal regions." Axoplasmic Transport (Ed.
D.G. Weiss), Springer-Verlag, Berlin.
[18] Jastrow, H. Electron Microscopic Atlas of Cells, Tissues and Organs
on the Internet. Schwann cells section.
http://www.uni-mainz.de/FB/Medizin/Anatomie/workshop/EM/EMSchwannE.html
[19] Kelly, F. R. Reversibility and stochastic networks. John Wiley & Sons
Ltd., Chichester, 1979. Wiley Series in Probability and Mathematical
Statistics.
[20] Lasek RJ, Brady ST (1982). The structural hypothesis of axonal trans-
port: two classes of moving elements. Axoplasmic Transport (ed. G.
Weiss), Springer-Verlag, Berlin, 1982, pp. 397-405.
[21] Lawler G (1995) Introduction to Stochastic Processes Chapman and
Hall, New York.
[22] Miller R, Lasek RJ (1985) Crossbridges mediate anterograde and ret-
rograde vesicle transport along microtubules in squid axoplasm. J Cell
Biol 101, 2181-2193.
[23] Moran, D., Rowley J.C. III Visual Histology.com, Chapter 8, Nerves.
http://www.visualhistology.com/products/atlas/VHA Chpt8 Nerves.html
[24] Ochs S (1972) Rate of fast axoplasmic transport in mammalian nerve
fibers. J Physiol 227, 627-245.
26
[25] Reed, M. C. and Blum, J. J. (1986) Theoretical analysis of radioactivity
proles during fast axonal transport: effects of deposition and turnover.
Cell Motility and the Cytoskeleton, 6, 620627.
[26] Reed M, Blum J (1994) Mathematical Questions in Axonal Transport.
In Lectures in mathematics in the Life Sciences,24, Amer. Math. Soc.,
Providence.
[27] Reed, M. C., Venakides, S. and Blum, J. J. (1990) Approximate travel-
ing waves in linear reaction-hyperbolic equations. SIAM J. Appl. Math.,
50, 167180.
[28] Schilstra MJ and Martin SR (2006) An elastically tethered viscous load
imposes a regular gait on the motion of myosin-V. Simulation of the
effect of transient force relaxation on a stochastic process. J.R. Soc. In-
terface 3 153-165.
[29] Schnitzer MJ, Visscher K and Block SM (2000) Force production by
single kinesin motors. Nat. Cell. Biol. 2 718-722.
[30] Stewart GH, Horwitz B, Gross GW. A chromatographic model of ax-
oplasmic transport. In Axoplasmic Transport (ed. G. Weiss), Springer-
Verlag, Berlin, 1982, 414-422.
[31] Svoboda K, Schmidt C, Schnapp B (1985) Direct observation of kinesin
stepping by optical trapping interferometry. Nature 365, 721-727.
[32] Takenaka T, Gotoh H (1984). Simulation of axoplasmic transport. J
Theor Biol 107, 579-601.
[33] Twiss JL, Fainzilber M (2009) Ribosomes in axons - scrounging from
the neighbors. Trends in cell biology. 19, 236-243
[34] Vale RD, Reese TS, Sheetz MP (1985). Identification of a novel force-
generating protein, kinesin, involved in microtubule-based motility. Cell
42, 39-50.
[35] Visscher K, Schnitzer MJ, Block SM (1999) Single kinesin molecules
studied with a molecular force clamp Nature 400, 184-189.
[36] Willis, WD, Grossman, RG (1981) Medical Neurobiology: Neu-
roanatomical and Neurophysiological Principles Basic to Clinical Neu-
roscience. Mosby, Inc.
1
1
0
2
y
a
M
0
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
2
2
7
2
.
1
1
9
0
:
v
i
X
r
a
A STOCHASTIC COMPARTMENTAL MODEL FOR FAST AXONAL
TRANSPORT
LEA POPOVIC∗, SCOTT A. MCKINLEY†, AND MICHAEL C. REED‡
Abstract. In this paper we develop a probabilistic micro-scale compartmental model and use
it to study macro-scale properties of axonal transport, the process by which intracellular cargo
is moved in the axons of neurons. By directly modeling the smallest scale interactions, we can
use recent microscopic experimental observations to infer all the parameters of the model. Then,
using techniques from probability theory, we compute asymptotic limits of the stochastic behavior
of individual motor-cargo complexes, while also characterizing both equilibrium and non-equilibrium
ensemble behavior. We use these results in order to investigate three important biological questions:
(1) How homogeneous are axons at stochastic equilibrium? (2) How quickly can axons return to
stochastic equilibrium after large local perturbations? (3) How is our understanding of delivery time
to a depleted target region changed by taking the whole cell point-of-view?
1. Introduction. In all cells, one finds that proteins, membrane-bound or-
ganelles, and other structures (e.g. chromosomes) are transported from place to place
at speeds much higher than diffusion. Though these transport processes are funda-
mental to cell function, many of the underlying mechanisms, organizational principles,
and regulatory features remain unknown. Axonal transport is one of the best studied
systems because the transport is basically one-dimensional since axons are long and
narrow. There are two speeds of axonal transport. Fast transport goes at speeds of
roughly 0.2 to 0.5 meters/day [27][33], while slow transport goes at approximately 1
millimeter/day, the rate of axon growth and regeneration [6][27]. The biology and
principles of slow transport are not yet clear [6], but the basic mechanisms of fast
axonal transport were discovered in the 1980s [2][3][29][43]. The model in this paper
refers to fast axonal transport, which we will henceforth call axonal transport.
The axonal transport apparatus consists of vesicles which form reversible chemical
bonds with motor proteins that bind reversibly to microtubules which run parallel to
the long dimension of the axon [1]. When the vesicle-motor protein complex is assem-
bled on the microtubule, the complex steps stochastically with step size approximately
8 nanometers for kinesin and dynein and 10 nanometers for myosin [7][12][18][40]. The
vesicles enter from the cell body on microtubules and then detach and reattach to the
transport mechanism at random times.
In this paper we propose a spatial Markov-chain compartmental model based on
these dynamics. We will assume independence of the interactions, and exponential
wait times between events. While we address the validity of these assumptions in the
Discussion section, we consider this a useful "first-order" approximation that permits
study of the dynamics from both the perspective of individual vesicles as well as that
of the full spatial system. Such a model unifies all earlier deterministic and stochastic
modeling efforts and can accommodate both qualitative and quantitative experimental
data observed on multiple scales.
In much experimental work in the 1970s and 1980s, radio-labeled amino acids
were put into the cell bodies continuously or for a few hours. The amino acids were
incorporated into proteins that were packaged into vesicles and put on the transport
system so that at later times radioactivity could be seen moving progressively down
∗Department of Mathematics and Statistics, Concordia University, Montreal QC H3G 1M8
†Department of Mathematics, University of Florida, 358 Little Hall Box 118105, Gainesville, FL
32605 ([email protected])
‡Department of Mathematics, Duke University, Box 90320, Durham, NC 27701
1
2
the axons.
In the continuous infusion case, one would see a wave of radioactivity
with a sharp but slowly spreading wavefront propagating at constant velocity down
the axon.
In the case of infusion for a few hours one would see at long times a
slowly spreading pulse of radioactivity that looked normally distributed. It was to
understand this behavior that Reed and Blum constructed PDE models for axonal
transport [3][34][35]. These models did not have traveling wave solutions, but the
data certainly looked like approximate traveling waves.
In [36] it was shown by a
perturbation theory argument that, in the asymptotic limit where the unbinding and
binding rates k2 and k1 get large, the solution approaches a slowly spreading traveling
wave or a normal pulse. Recently, in a series of papers, Friedman and co-workers have
introduced new PDE models and proved these results rigorously[14][15][16][17].
Probabilistic models for axonal transport were introduced and used for simula-
tions already in the 1980s [39][41]. However, rigorous work began with Lawler [28] in
1995 and was continued by Brooks [5] who used a continuous time stochastic model
to show that the distribution of an individual particle is a spreading Gaussian at large
times. Brooks also proved tail estimates for the central limit theorem and used them
to estimate the error from normal. Independently, Bressloff [4] developed a discrete
stepping model and performed an analysis under the assumption that the rate of un-
binding and binding to transport is fast relative to lateral velocity over the length
scale of interest. The author derived a characterization of the spreading wavefront of
a particle entering at the nucleus and traveling to the distal end. This model served as
the basis for later investigations by Newby and Bressloff [31, 32] wherein the authors
characterize the axonal transport system as an intermittent search for hidden targets.
In this paper we revise the existing probabilistic models in order to study ran-
domness in the system as a whole rather than exclusively from the point of view of
an individual particle. Our goal is not only to recover and generalize previous results,
but also to investigate three specific, biologically important, properties of the whole
stochastic system.
1.1. Summary of Results. In Section 2, we create a continuous-time Markov
chain queueing model for the axonal transport system. We show how to use experi-
mental data to determine (or estimate) all the parameters of the model.
In Section 3, we take the individual vesicle point-of-view. We prove the asymp-
totic forms in [36] with rigorous error estimates. We show that in the limit as the
compartment size becomes small our model becomes the probabilistic model of [5].
We also show that in the limit as the length of the axon and time become large (with
the scale of axon length on the order of the squared scale of time) our model becomes
the PDE model of [35]. Since we assume that particles are independent, the time evo-
lution of the law of an individual will reflect the behavior of an ensemble of particles
released at the same time.
In Section 4 we adopt the full spatial system perspective to quantify stochasticity
along the length of the axon. We begin by calculating in Proposition 4.1 the stationary
distribution of a flow-through system that has sustained input from the nucleus, while
particles are removed upon reaching the distal end. The stationary distribution has a
product Poisson structure which allows for seamless transition between spatial scales.
With this mathematical model, we are able to make precise statements about
three biologically important system properties.
In the stationary distribution, the
number of vesicles in each slice of the axon is independent and identically distributed
(Proposition 4.1), however this in and of itself is not sufficient to account for the
sense that samples taken for different parts of the cell "look the same." We compute in
3
Section 4.2 the coefficient of variation for the number of vesicles in sections of different
length and show that the coefficient of variation is low for all but the smallest length
scales.
In Section 4.3, we study the intermittent search problem posed by Newby
and Bressloff [31] from the system point of view. Efficient transport to locations that
need material must balance the speed of transport of material from the nucleus to the
distal end of the cell with the rates of dissociation from the transport apparatus. We
calculate the expected hitting time for a hidden target by all vesicles in the system.
In so doing we encounter the counterintuitive result that while increasing the velocity
of the motors while on transport increases the chance of any particular vesicle missing
the target, the expected hitting time by the system actually decreases.
This hitting time approach is natural for needed material that is sparsely dis-
tributed throughout the axon, but when the needed cargo in question is more com-
mon, the time to replenishment is better addressed through the ODE approach that
we develop in Section 4.4. Due to the product structure of the law of the transient
dynamics, this non-equilibrium behavior is determined by the 2N -dimensional ODE
governing the means. From this we estimate the timescale of return to equilibrium as
a function of the length scale of interest.
2. The model and its parameters. Let L be the length of the axon, divided
evenly into N = L/δ lateral sections each of length δ, equal to the step size of the
motor protein. Within each section, we disregard any further spatial geometry and
take the particles to be in one of two states:
· an on-transport state that steps laterally at a rate r = v/δ per section, or
· an off-transport state that does not step laterally.
We use a 2N -dimensional continuous time Markov chain {(Qi(t), Pi(t)), i = 1, . . . , N}
to model the particle dynamics, where
· Qi(t) is the number of particles at time t in on-transport state in section i,
· Pi(t) is the number of particles at time t in off-transport state in section i.
be a continuous time Markov chain on the state space NN×NN with the following tran-
sitions and time dependent rates:
Definition 2.1. (Stochastic compartmental model) Let ({(Qi(t), Pi(t)), i = 1, . . . , N})t≥0
· (Lateral transport)
(Qi, Qi+1) → (Qi − 1, Qi+1 + 1) at rate rQi(t);
· (Switch from on-transport to off-transport)
(Qi, Pi) → (Qi − 1, Pi + 1) at rate k2Qi(t);
· (Switch from off-transport to on-transport)
(Pi, Qi) → (Pi − 1, Qi + 1) at rate k1Pi(t);
· (Production of new particles)
(Q1) → (Q1 + 1) at rate rq0;
· (Removal of particles at distal end)
(QN ) → (QN − 1) at rate rQN (t).
The lateral transport rate, r = v/δ, is inversely proportional to the length scale
so that the mean number of particles per unit length is invariant with respect to
rescaling δ. We will assume that the rate of production q0 = δρ0, for some constant
ρ0 > 0, in order for the mean number of particles in each compartment to scale with
the size δ of each compartment. This will imply that the mean number of particles
per unit length scales as ρ0. A graph of the model is depicted in Figure 1.
In order to insure the Markov property, we use exponential random variables for
the waiting times between transition events. Specifically, we mean that after a given
4
Q1 = 3
P1 = 3
∗
•
∗
∗
∗
•
x = L
3
Q2 = 1
P2 = 2
∗
•
•
N
u
c
l
e
u
s
x = 0
Q3 = 3
P3 = 2
•
∗
∗
•
•
x = 2L
3
x = L
Q1
✲q0r
Q2
✲r
Q3
✲r
✲r
❑
k1
k2
❯
P1
❑
k1
k2
❯
P2
❑
k1
k2
❯
P3
Fig. 2.1. On-and-off Transport Chain
event we assign a new independent random variable to each of the 3N +2 possible next
events, exponentially distributed with the appropriate rate parameter. The system
of values updates according to the transition associated with the minimum of these
waiting times. Then we create a new set of exponential random variables and the
process proceeds as before.
The advantage of computing explicit formulas for quantities that can be observed
in experiments is that the experimental data can then be used to determine the
parameter values in the model. For the characterization of the approximate wavefront
speed and spreading in Section 3 and the homogeneity calculations in Section 4 we
need order-of-magnitude estimates for the parameters. Actual parameter values will
certainly differ depending on the particular neural tissue and the particular particles
being transported. However, we can get order-of-magnitude estimates from existing
data.
First we recall that fast transport has been observed to travel at speeds of 0.2 to
0.5 m per day. We can assume that the average velocity of particles while physically
bound to microtubules is roughly 1 m per day, or v = 10−6 m/s. We assume that the
compartment size scales as the length scale of the individual steps of the motor protein,
so δ ∼ 10−8m. This implies that the rate parameter should be r = v/δ = 100s−1.
We now turn our attention to the on-off rates rates k2 and k1. These can be de-
termined from experimentally observed run lengths on the transport system. Indeed,
5
r
r+k2
and therefore the average run length is r
Dixit et al. [10] show that a typical run along microtubules for dinein and kinesin
is on the order of 10−6m. We can compare this with the theoretical run length of
the model to determine off-rate k2. Within the model, at each step on the transport
mechanism the particle has a binary decision to jump laterally along the transport
with probability r/(k2 + r), or to jump off with probability k2/(r + k2). The number
of jumps along the transport system before jumping off is therefore geometrically dis-
tributed on the set {0, 1, . . .} with success probability
. It follows that average
number of steps in the run is r
k2 × 10−8 m.
k2
Setting this equal to the average experimental run length of 10−6 from [10], we see
that r
k2 ∼ 100, implying that k2 ∼ 1s−1. As we will see in the computation of the
stationary distribution in Section 4.1, the ratio of the expected number of particles
on the track to those off the track is k2
. Dixit [10] found that approximately 75% of
k1
the particles were motile so this ratio is approximately equal to 3. Since k2 ∼ 1s−1
we see that k1 ∼ 1
3 .
It remains to estimate q0. We will see in Proposition 4.1 that the mean number
of particles per compartment is (1 + k2
)q0 = 4q0. Of course, axons have a large
k1
variety of diameters and larger axons will have more vesicles per unit length so one
expects a range of values for q0. However, examination of a large number of electron
micrographs of axonal cross-sections (see for example [20], Fig. 3; [21]; [30]), which
are typically 100 nm thick enables one to estimate the number of vesicles per 100 nm
segment. This number is typically in the range of 10 to 100 which implies that there
are 1 to 10 vesicles per compartment in our model. Therefore q0 is in the range 0.25
to 2.5, for various axons.
We remark that we are ignoring some aspects of the physics and the biology of
axonal transport. We are not including diffusion of the vesicles off the track. We
are treating the microtubule track as though it were a single continuous entity from
one end of the axon to the other, when it fact it consists of numerous, separated,
microtubule fragments. And, we are ignoring retrograde transport and the details of
the motor proteins. Nevertheless, this simple model will enable us to investigate the
homogeneity questions that are the main goal of this paper.
3. Dynamics from the Particle Perspective. In this section we calculate
properties of the stochastic dynamics by using stochastic convergence theorems and
stochastic averaging theorems from probability theory. We first see that, in the δ → 0
limit, the law of the location of a single particle corresponds to that of a particle with
a piecewise linear Markov motion. We then show this law can be approximated by the
Green's function of a linear partial differential equation. This enables us to obtain,
as a special case, the asymptotic behavior of the PDE models for axonal transport in
an either rate limiting or perturbed setting.
3.1. The active transport mode. We first consider the simple case where
0 = 0 and stays exclusively in active transport mode. Let
t be the
the particle starts at X δ
X δ
t ∈ {0, δ, 2δ, . . . , L} be the lateral position of a particle at time t and let nδ
number of jumps made by the particle as of time t. Observe, X δ = δnδ.
Lemma 3.1. Let k2 = k1 = 0, and r = v/δ > 0, then the position of the particle
satisfies, for any time t < ∞
sup
s≤t(cid:12)(cid:12)X δ
s − vs(cid:12)(cid:12) −→δ→0
0 a.s.
6
and, for B a standard Brownian motion
1
√δ(cid:16)X δ
t − vt(cid:17)t≥0
=⇒δ→0
√v(cid:0)Bt(cid:1)t≥0
in distribution on the Skorokhod space of cadlag (right continuous left limited) func-
tions.
Proof. Since nδ is a Poisson process with rate r = v/δ, defining Nt := nδ
δt we get
a Poisson process N with rate v, and we have X δ
· = δN·/δ. Our results then follow
directly from the functional law of large numbers (FLLN) and the functional central
limit theorem (FCLT) for the Poisson process N .
3.2. The on/off dynamics. We now consider a particle which undergoes tran-
sitions from on-transport to off-transport state and back. Denote again by X δ
t ∈
{0, δ, 2δ, . . . , L} the lateral position of a particle at time t and let nδ
t be the number of
lateral transition jumps made by the particle as of time t. Observe that the particle
will spend only a fraction of its time in active transport and hence the lateral speed
of the particle should be slower than before.
A non-compartmental stochastic model for axonal transport, introduced by Brooks
in [5], is as follows. A particle can be in one of two states:
We use a 2-dimensional Markov process to model the particle dynamics, where
· an on-transport state with deterministic lateral velocity v, or
· an off-transport state with lateral velocity 0.
· Xt be the lateral position of this particle at time t,
· ξt be the indicator for whether it is on (1) or off (0) transport at time t.
Definition 3.2.
(Stochastic non-compartmental model) Let (Xt, ξt)t≥0 be a
piecewise-linear Markov process with values in (R+,{0, 1}) started at (X0, ξ0) = (0, 1)
with the following dynamics:
· (Switch from on-transport to off-transport)
(Xt, 1) → (Xt, 0) at rate k1
· (Switch from off-transport to on-transport)
(Xt, 0) → (Xt, 1) rate k2
· (Lateral travel) Xt =R t
0 vξsds
The path of (Xt)t≥0 consists of alternating sequence of Exponential(k2) stretches of
time where the lateral position increases linearly with speed v, and Exponential(k1)
stretches of time where it remains constant.
Proposition 3.3. Let k2, k1 > 0, and r = v/δ > 0, then the position of the
particle converges
(X δ
t )t≥0 =⇒δ→0
(Xt)t≥0
in distribution on the Skorokhod space of cadlag functions.
Proof. If we let ξδ be the indicator of whether the particle in the compartmental
model is on (ξδ = 1) or off (ξδ = 0) transport, then (X δ
t )t≥0 is a strong Markov
process. We will see that ξδ is a continuous-time Markov chain on {0, 1} (independent
of δ), and conditionally on ξδ the transition law of X is easily expressed. Likewise,
for the non-compartmental model above, (Xt, ξt)t≥0 is a strong Markov process, with
ξ the same continuous-time Markov chain on {0, 1} as ξδ, and conditionally on ξ the
change in X is easily given in terms of its linear speed and ξ.
t ) converges to (Xt, ξt) in
distribution as δ → 0. We then show that the finite dimensional distributions of
We will start by showing that for any t > 0, (X δ
t , ξδ
t , ξδ
(X δ, ξδ) converge to those of (X, ξ). A tightness argument finally implies (Lemma
16.2 and Theorem 16.3 [22]) that (X δ, ξδ) converges to (X, ξ) in distribution on the
Skorokhod space of cadlag processes.
7
0 , ξδ
Suppose that initially the particle is on transport at x0, so (X δ
0) = (x0, 1).
The first time τ1 = inf{t > 0 : ξδ
t = 0} at which the particle steps off transport
has Exponential(k2) distribution, irrespective of δ. The first subsequent increment in
time σ1 = inf{t > 0 : ξδ
τ1+t = 0} after which the particle steps back on transport has
Exponential(k1) distribution, irrespective of δ as well. This is repeated, and ξδ is a
simple continuous-time Markov chain on {0, 1} with transition rates k2 and k1, from
1 → 0 and 0 → 1, respectively.
Until time τ1 the particle behaves as if it were in active transport (k2 = k1 = 0)
and conditionally on the value of τ1, for any 0 ≤ s ≤ τ1 the lateral change in position
over time s, X δ
s is a Poisson(rs) random variable. Moreover,
by results of Proposition 3.1, conditionally on the value of τ1, we have
s where nδ
s − X δ
0 , is δnδ
0 a.s.
sup
0≤s≤τ1(cid:12)(cid:12)(X δ
s − X δ
0 ) − vs(cid:12)(cid:12) −→δ→0
τ1 − X δ
To get the unconditioned law of X δ
0 , we observe that the number of boxes
the particle traverses before it steps off nδ
τ1 has a Geometric(k2/(k2 + r)) distribution
(note that a Poisson rate r process sampled at an Exponential(k2) time independent
of the process has this distribution). Since k2/δ(k2 + r) → k2/v, as δ → 0, X δ
0 =
δnδ
τ1 converges in distribution to Exponential(k2/v) variable.
τ1 −X δ
Consider now the particle in the non-compartmental model. It is immediate from
the definition of the model that ξ has the same law of a continuous-time Markov chain
on {0, 1} with transition rates k2 and k1, from 1 → 0 and 0 → 1, as ξδ. Since we prove
our convergence in law results by first conditioning on the values of ξδ and ξ, we can
without loss of generality henceforth assume ξδ = ξ a.s., and drop its superscript.
Suppose that initially the particle in the non-compartmental model is on transport
at x0, so (X0, ξ0) = (x0, 1). Conditionally on τ1, for all 0 ≤ s ≤ τ1, Xs − X0 = vs and
Xτ1 − X0 = vτ1, hence unconditionally, Xτ1 − X0 is an Exponential(k2/v) random
variable. We now have both sup0≤s≤τ1(cid:12)(cid:12)(X δ
τ1 − X δ
X δ
ally on values of τ1, σ1, and of X δ
τ1+σ1−X δ
and X δ
tial values (X δ
respectively.
0 ) − (Xs − X0)(cid:12)(cid:12) → 0 a.s. and
τ1)− (Xs − Xτ1)(cid:12)(cid:12) ≡ 0,
0 ⇒ Xτ1+σ1−X0. At time τ1+σ1, the same process starts over from ini-
τ1, 1) and (Xτ1 , 1) in the compartmental and non-compartmental model,
τ1, Xτ1, supτ1≤s≤τ1+σ1(cid:12)(cid:12)(X δ
Between times τ1 and σ1 the particle in both models stays in place, so condition-
0 ⇒ Xτ1 − X0.
s − X δ
s − X δ
Let τ1, σ1, τ2, . . . , be the sequence of time increments between the consecutive
times when the particle in both models gets off and gets back on transport, let σ0 = 0
and for i ≥ 1
τi = inf{t > 0 : ξσi−1 +t = 0},
σi = inf{t > 0 : ξτi+t = 1}
Then (τi)i≥1 and (σi)i≥1 are independent sequences of i.i.d. Exponential(k2) and
Exponential(k1) variables, respectively. For any t > 0, let ηt be the number of times
the particle in either model gets back on transport until time t, and η′
t the number of
times it gets off,
ηt = inf{k ≥ 0 :
k
Xi=1
(τi + σi) ≤ t},
t = inf{k′ ≥ 0 :
η′
τi +
k′
Xi=1
k′−1
Xi=1
σi ≤ t}
8
t = ηt iff ξt = 1, and η′
Note that, η′
t = ηt + 1 iff ξt = 0. Let τt be the last time
before time t that the particle changed whether it was on or off transport, that is,
τt = sup{0 ≤ s ≤ t : ξs− 6= ξs}. Then, we have
τt =(
ηt
η′
t
Pi=1
Pi=1
ηt
Pi=1
(τi + σi)
if η′
t = ηt,
τi +
σi
if η′
t = ηt + 1.
If η′
t = ηt, then from time τt to t the particle is in active transport, and the same
convergence argument as before implies that conditionally on the values of τt and X δ
τt,
τt≤s≤t(X δ
sup
s − X δ
τt) − v(s − τt) → 0 a.s.
If η′
0 = δnδ
τt − X δ
τt − X δ
τt where nδ
Also X δ
τt is the number of boxes the particle traverses by time
t. Conditionally on the value of ηt, nδ
τt is a sum of ηt i.i.d. Geometric(k2/(k2 + r))
random variables, hence X δ
0 converges in distribution to a sum of ηt i.i.d.
Exponential(k2/v) random variables. In the non-compartmental model, if η′
t = ηt,
then conditionally on the values of ηt and τt, for τt ≤ s ≤ t, Xs − Xτt = v(s− τt) and
conditionally only on the value of ηt, X δ
τt − X0 is a sum of ηt i.i.d. Exponential(k2/v)
variables. Hence, conditionally on ηt and τt, supτt≤s≤t(cid:12)(cid:12)(X δ
τt − X δ
so conditionally on ηt and τt, supτt≤s≤t(cid:12)(cid:12)(X δ
0 ⇒ Xτt − X0.
s − X δ
τt is a sum of η′
τt)− (Xs − Xτt)(cid:12)(cid:12) → 0
τt)− (Xs − Xτt)(cid:12)(cid:12) = 0 a.s.. Also, con-
t = ηt + 1, then from time τt to t the particle is in both models stays in place,
ditionally only on the value of η′
ables, hence X δ
variables. In the non-compartmental model, if η′
value of η′
only on η′
t i.i.d. Geometric(k2/(k2 + r)) vari-
t i.i.d. Exponential(k2/v)
t = ηt + 1, conditionally only on the
t i.i.d. Exponential(k2/v) variables. Hence, conditionally
0 converges in distribution to a sum of η′
a.s., and conditionally only on ηt, X δ
τt −X δ
τt is a sum of η′
τt − X δ
Now, integrating over the possible values of ηt,η′
t and τt, we get that for any t ≥ 0,
(X δ
t , ξδ
t ) ⇒ (Xt, ξt). Convergence of finite dimensional distributions follows from an
iterative use of the Markov property of (X δ, ξδ) and (X, ξ), and the fact that the
increments of both (X δ, ξδ) and (X, ξ) are stationary.
0 ⇒ Xτt − X0.
t, X δ
t, X δ
s − X δ
t, nδ
In order to verify tightness, Theorem 16.11 [22], of the sequence of Markov pro-
cesses {(X δ, ξδ)}δ>0, because (X δ, ξδ) has stationary increments and is strong Markov,
it will suffice to check that for any ǫ > 0
lim
h→0
lim sup
δ→0
h, ξδ
h) − (X δ
0 , ξδ
P(cid:8)(X δ
0) > ǫ(cid:9) = 0
where (x1, ξ1)−(x2, ξ2) = x1−x2+ξ1−ξ2 is a distance metric on R+×{0, 1}. For
any δ > 0, the first change in the continuous-time Markov chain ξδ happens after an
Exponential(k) time (where k = k2 or k = k1 depending on whether ξδ
0 = 0),
0 = 1 or ξδ
0(cid:9) ≤ 1 − e−kh. Moreover,
h 6= ξδ
and is independent of δ. Hence, at time h later, P(cid:8)ξδ
h− X δ
irrespective of the value of ξδ, at time h later the value of X δ
0 ≤ δnδ
0 > ǫ(cid:9) ≤ δE(cid:2)nδ
h − X δ
a Poisson process with rate r = v/δ. Hence, P(cid:8)X δ
h(cid:3)/ǫ = vh/ǫ.
Combining the two gives
0) > ǫ(cid:9) ≤ 1 − e−kh + vh/ǫ for any δ > 0,
h) − (X δ
and the desired limit follows.
P(cid:8)(X δ
h where nδ is
h, ξδ
0 , ξδ
9
t ) ∈ δZ+ × {0, 1} is a Markov process
The process (X δ, ξδ) : t ∈ [0,∞) 7→ (X δ
t , ξδ
with cadlag paths whose generator is given by
Aδf (x, ξ) = rξ(cid:2)f (x + δ, ξ) − f (x, ξ)(cid:3)
+ k2ξ(cid:2)f (x, ξ − 1) − f (x, ξ)(cid:3) + k1(1 − ξ)(cid:2)f (x, ξ + 1) − f (x, ξ)(cid:3)
for all f ∈ D(Aδ) = C0(δZ+ × {0, 1}).
The piecewise linear process (X, ξ) : t ∈ [0,∞) 7→ (Xt, ξt) ∈ R+ × {0, 1} is a
Markov process with continuous paths whose generator is the closure of the operator
Af (x, ξ) = vξ∂xf (x, ξ)
+ k2ξ(cid:2)f (x, ξ − 1) − f (x, ξ)(cid:3) + k1(1 − ξ)(cid:2)f (x, ξ + 1) − f (x, ξ)(cid:3)
for all f ∈ D(Aδ) = C1,0(R+ × {0, 1}).
Letting ιδ : δZ+ × {0, 1} 7→ R+ × {0, 1} be an embedding, and f δ = f ◦ ιδ, then
Aδf δ → Af as δ → 0 for all f ∈ C1,0(R+ × {0, 1}) imply that the finite dimensional
distributions of (X δ, ξδ) converge to those of (X, ξ). Verification of additional con-
ditions, see Theorem 19.25 of [22], would also imply convergence of processes with
generators {(cid:0)Aδ,D(Aδ)(cid:1)}δ>0 to the process with generator (cid:0)A,D(Aδ)(cid:1), however, we
thought this way of showing convergence in law was not as instructive.
The fact that an individual particle will have the distribution given by Propo-
sition 3.3 as the size of the boxes decreases means that our model is a microscopic
version of the stochastic model used by Brooks [5], and that the hydrodynamic limit
of our model as δ → 0 is equal to the macroscopic stochastic model from [5].
An approximation of the particle's position Xt is obtained in [5] to be Xt ≈
µt + √tσZ as t → ∞, where µ = k1v/(k2 + k1), σ = 2k2k1v2/(k2 + k1)3 and Z is
a standard Normal variable. That approximation is valid only for large fixed values
of t, while we next extend this result to give an approximation for the whole time
trajectory of the particle's path. This is accomplished by the following functional
central limit theorem for the position of the particle undergoing stochastic transport.
Proposition 3.4. Let X be the position of a particle following the piecewise
linear Markov process from Proposition 3.3 started at X0 = 0 on transport, then
and, if B denotes a standard Brownian motion
Xns
n −
k1
k2 + k1
sup
s≤t(cid:12)(cid:12)
√n(cid:16) Xnt
n −
0 a.s. ∀t > 0
vs(cid:12)(cid:12) −→n→∞
=⇒n→∞ s 2k2k1v2
k1v
k2 + k1
t(cid:17)t≥0
(k2 + k1)3(cid:0)Bt(cid:1)t≥0
in distribution on the space of continuous functions.
Proof. For these results we use the notion of stochastic averaging [23][24][26].
Note that the indicator process ξ for being on- or off- transport is independent of the
position X of the particle. Hence, the position of the particle X is a linear random
evolution process, Ch 12 of [11], [19], driven by the independent indicator process ξ.
The generator of (X, ξ) is the closure of the operator
Af (x, ξ) = σ(ξ)∂xf (x, ξ) + λ(ξ)(cid:2)f (x, s(ξ)) − f (x, ξ)(cid:3)
10
for all f ∈ D(A) = C1,0
functions in x continuous in ξ and vanishing at infinity), where
0 (R+ × {0, 1}) (the space of all continuously differentiable
σ(ξ) = vξ, λ(ξ) = k2ξ + k1(1 − ξ), and s(ξ) = 1 − ξ
(when ξ = 1: σ = v, λ = k2, s = 0 and when ξ = 0: σ = 0, λ = k1, s = 1). In other
words, if (X0, ξ0) = (0, 1) we have
Xt =Z t
0
vξsds,
ξt =
1
2(cid:0)1 + (−1)Y (R t
0 λ(ξs)ds)(cid:1)
0 λ(ξs)ds) is a counting process of the
where Y is a rate 1 Poisson process, and Y (R t
number of switches of ξ until time t.
Rescaling time and position of the process by 1/n, we get that ( Xn·
n , ξn·) satisfies
Xnt
n
=Z t
0
vξnsds,
ξnt =
and its generator is
1
2(cid:0)1 + (−1)Y (n R t
0 λ(ξns)ds)(cid:1)
Anf (x, ξ) = vξ∂xf (x, ξ) + n(k2ξ + (k1(1 − ξ)))(cid:2)f (x, s(ξ)) − f (x, ξ)(cid:3)
for all f ∈ D(An) = C1,0
0 (R+ × {0, 1}). Note that ξn· switches at rate propor-
tional to n, forming an ergodic Markov chain with stationary distribution π(1) =
k1/(k2 + k1), π(0) = k2/(k2 + k1), andR σ(ξ)π(ξ) = vR ξπ(ξ) = vk1/(k2 + k1). Hence,
the strong ergodic theorem implies that
Xn
n
=
1
nZ n
0
vξsds −→n→∞
vk1
k2 + k1
a.s.
We can extend this to a functional statement on any finite time interval [0, t]. Fix
t > 0 and take any ∆ > 0, then there exists n∆ < ∞ a.s. such that
Xn
n −
(cid:12)(cid:12)
vk1
k2 + k1(cid:12)(cid:12) <
∆
t
,
for ∀n > n∆
Now, let M = supn≤n∆ Xn − k1/(k2 + k1)vn, which is finite a.s. since n∆ < ∞ a.s.
Let n > M/∆. Then, for any 0 ≤ s ≤ t we have that either ns > n∆ in which case
vk1
∆
t
vk1
k2 + k1
(cid:12)(cid:12)
(cid:12)(cid:12)
ns −
s ≤ ∆
Xns
n −
or, ns ≤ n∆ in which case
s(cid:12)(cid:12) =(cid:12)(cid:12)(cid:16) Xns
s(cid:12)(cid:12) =
n(cid:12)(cid:12)Xns −
n Xns − vk1/(k2 + k1)s(cid:12)(cid:12) < ∆ whenever n > M/∆.
implying that we have sup0≤s≤t(cid:12)(cid:12)
Once we rescale the position for the particle by 1/√n and time by 1/n, ξ still
changes at a much faster rate than the position of the particle X. The generator of
the rescaled centered process (X n, ξn) defined as
k2 + k1(cid:17)s(cid:12)(cid:12) <
ns(cid:12)(cid:12) <
Xns
n −
k2 + k1
k2 + k1
M
n
< ∆
vk1
1
vk1
1
X n
t := √n(cid:0)
Xnt
n −
k1v
k2 + k1
t(cid:1),
ξn
t := ξnt
is the closure of the operator
11
¯Anf (x, ξ) =(cid:0)√nσ(ξ) −
k1v
k2 + k1(cid:1)∂xf (x, ξ) + nλ(ξ)(cid:0)f (x, s(ξ)) − f (x, ξ)(cid:1)
on f ∈ D(An) = C1,0
0 (R × {0, 1}). We will use the stochastic averaging theorem
(Theorem 2.1 of [26]) to show that the paths of the centered rescaled process converge
in distribution to paths of a Brownian motion with a diffusion coefficient equal to
2k2k1/(k2 + k3)3v2.
Let h(ξ) be the function
h(ξ) = v
k2k1
1
(k2 + k1)3
λ(ξ)
= v
k2k1
1
(k2 + k1)3
k2ξ + k1(1 − ξ)
Then h(1) = vk1/(k2 + k1)2, h(0) = −vk2/(k2 + k1)2 imply that h(s(1)) − h(1) =
−v/(k2 + k1), h(s(0)) − h(0) = v/(k2 + k1), which in turn imply that λ(1)(cid:0)h(s(1)) −
h(1)(cid:1) = −vk2/(k2 + k1), λ(0)(cid:0)h(s(0)) − h(0)(cid:1) = vk1/(k2 + k1), so that for ξ ∈ {0, 1}
k1
λ(ξ)(cid:0)h(s(ξ)) − h(ξ)(cid:1) = −(cid:0)vξ − v
k2 + k1(cid:1)
0 (R) define a sequence of functions f n ∈ C1,0
0 (R × {0, 1}) by
Now, for any f ∈ C2
f n(x, ξ) = f (x) +
1
√n
h(ξ)∂xf (x)
Then f n → f as n → ∞ and
¯Anf n(x, ξ) = √n(cid:0)vξ − v
k2 + k1(cid:1)∂xf (x) +(cid:0)vξ − v
+ √nλ(ξ)(cid:0)h(s(ξ)) − h(ξ)(cid:1)∂xf (x)
k2 + k1(cid:1)h(ξ)∂2
=(cid:0)vξ − v
xf (x) = ¯Af (x)
k1
k1
where ¯A is defined on D( ¯A) = C2
0(R) by
k1
k2 + k1(cid:1)h(ξ)∂2
xf (x)
¯Af (x) =
k2k1v2
(k2 + k1)3 ∂2
xf (x)
Define a sequence of processes
εf,n
t =
1
√n
h(ξn
t )∂xf (X n
t ) = f n(X n
t , ξn
t ) − f (X n
t )
Then our earlier calculation implies that for any f ∈ D( ¯A)
t ) −Z t
t ∈ {0, 1} it is clear that
is a sequence of martingales. Since f ∈ C2
t ) −Z t
0 (R), ξn
s )ds + εf,n
t = f n(X n
Anf (X n
¯Af (X n
f (X n
t , ξn
0
0
s , ξn
s )ds
sup
n
E(cid:20)Z t
0 (cid:12)(cid:12)
¯Af (X n
2
ds(cid:21) < ∞ and E(cid:20)sup
s
s≤t(cid:12)(cid:12)εf,n
0
(cid:12)(cid:12)(cid:21) −→n→∞
s )(cid:12)(cid:12)
12
In order to apply Theorem 2.1 of [26] on stochastic averaging it is only left to show
the process X n satisfies the compact containment condition, that is for any t > 0 and
∆ > 0 there exists a compact set K ⊂ R such that
P{X n
This follows from the fact that X n
f (x) = x) with mean
inf
n
s ∈ K ∀s ≤ t} ≥ 1 − ∆
t + h(ξn
t )/√n is a sequence of martingales (let
E(cid:20)X n
t +
h(ξn
t )
√n (cid:21) = X n
0 +
h(ξn
0 )
√n
=
k2k1
(k2 + k1)3
v2
√n
and second moment (let f (x) = x2)
h(ξn
t )
√n (cid:17)2(cid:21) =
k2k1
(k2 + k1)3 v22t +
E[h(ξn
t )]
n
So, by Doob's inequality
t +
E(cid:20)(cid:16)X n
√n (cid:12)(cid:12)(cid:12) ≥ C(cid:27) ≤
h(ξn
t )
E[h(ξn
t )]
(cid:17)
4
h(ξn
t )
t +
k2k1
X n
s +
4
C2
C2(cid:16)
E(cid:20)(cid:16)X n
(k2 + k1)3 v22t+
√n (cid:17)2(cid:21) =
P(cid:26)sup
s≤t(cid:12)(cid:12)(cid:12)
Noting that hmin := v min(k2, k1)/(k2 + k1)2 ≤ h ≤ hmax := v max(k2, k1)/(k2 + k1)2,
and choosing C (given on t and ∆) so that the right hand side of the inequality with
n = 1 is less than ∆, shows that with K = [−C − hmax, C − hmin] the compact
containment condition holds for (X n)n≥1.
Now Theorem 2.1 of [26] implies that X n ⇒ W in distribution on the Skorkhod
space of continuous functions, where W is a process with generator ¯A, and conse-
quently has the same distribution as p2k2k1v2/(k2 + k1)3B where B is a standard
Brownian motion.
n
3.3. Connection to Partial Differential Equations Models. In order to
demonstrate the connection between our model and the PDEs seen in [36][35][16],
consider the process (X, ξ) of the particle following the piecewise linear Markov pro-
cess, and for any x ≥ 0, t ≥ 0 let
q(x, t) = P{Xt ∈ dx, ξt = 1}/dx,
p(x, t) = P{Xt ∈ dx, ξt = 0}/dx
denote the probability densities of the particle's location x on and off transport,
respectively, over time. Kolmogorov forward equations for (X, ξ) imply that q and p
satisfy the system of PDEs
∂tq(x, t) + v∂xq(x, t) = −k2q(x, t) + k1p(x, t)
∂tp(x, t) = k2q(x, t) − k1p(x, t).
(3.1)
(3.2)
When k2 = 0 = k1, the limiting PDE is simple linear transport: (∂t+v∂x)q(x, t) =
0. The initial condition q(x, 0) = δ0(x) corresponds to the density of a single particle
at the origin at t = 0. The time evolution via simple linear transport is translation of
the delta function, while the time evolution via the equations (3.1) and (3.2) will have
a spreading profile. This is clear from the macroscopic limits of (X δ, ξδ) as δ → 0.
When k2 = k1 = 0 the particle never switches off from traveling on transport at speed
v and is deterministic, as seen in Proposition 3.1. When k2, k1 > 0 the particle follows
13
a truely stochastic process (X, ξ) with a non-zero variance, as seen in Proposition 3.3
and Proposition 3.4.
In the experiments described in the introduction one sees "approximate" traveling
waves of radioactivity in the axons in the sense that there is a slowly spreading wave
front moving at constant velocity away from the cell body. Equations (3.1) and (3.2)
are linear and do not have solutions that are bounded traveling waves. It was shown
by a perturbation theory argument in [36][35] that as ε → 0 the solutions of
ε(∂t + v∂x)qε(x, t) = −k2qε(x, t) + k1pε(x, t),
ε∂tpε(x, t) = k2qε(x, t) − k1pε(x, t).
(3.3)
(3.4)
subject to qε(0, t) = q0, are to leading order
qε(x, t) = c1H(
x − µt
ε1/2 , t),
pε(x, t) = c2H(
x − µt
ε1/2 , t),
where H satisfies the heat equation
∂sH(y, s) =
σ2
2
∂yyH(y, s),
H(y, 0) = χ(−∞,0)
(3.5)
µ =
k2v
k2 + k1
, σ2 =
2k2k1v2
(k2 + k1)3
c1 =
k1
k2 + k1
,
c2 =
k2
k2 + k1
.
This asymptotic form is valid for small ε, that is for large k2 and k1. However, if
we set q(x, t) = qε( x
ε ), then q and p satisfy (3.3) and (3.4),
so the solutions of (3.1) and (3.2) behave like approximate traveling waves for large
t and large x whether or not k2 and k1 are large. These results have been proven
rigorously by Friedman and coworkers [14][15][16][17].
ε ) and p(x, t) = pε( x
ε , t
ε , t
To see that our Proposition 3.4 provides another rigorous proof of these properties,
n Xn·, ξn·)
albeit using stochastic methods, consider the process (εX·/ε, ξ·/ε) (that is ( 1
with n = 1/ε), and let
qε(x, t) = P(cid:8)εXt/ε ∈ dx, ξt/ε = 1(cid:9)/dx,
pε(x, t) = P(cid:8)εXt/ε ∈ dx, ξt/ε = 0(cid:9)/dx
be the probability densities for this process. The generator of this process is An
(n = 1/ε), so the Kolmogorov forward equations imply that qε and pε satisfy the
system of PDEs (3.3) and (3.4). Our result from Proposition 3.4 states that
P(cid:8)εXt/ε ∈ dx(cid:9)/dx ≈ H(
x − µt
ε1/2 , t)
for small ε > 0, where H satisfies (3.5). Hence, qε+pε ≈ H( x−µt
ε1/2 , t), and P(cid:8)ξt/ε = 1(cid:9) ≈
k1/(k2 + k1), P(cid:8)ξt/ε = 0(cid:9) ≈ k2/(k2 + k1), gives the result that qε(x, t) and pε(x, t) are
well approximated by
ε1/2 , t) and
H( x−µt
H( x−µt
ε1/2 , t).
k1
k2+k1
k2
k2+k1
4. Dynamics from the spatial system perspective.
4.1. The spatial system in equilibrium. We are now ready to characterize
the steady state dynamics induced by continually adding particles from the nucleus
and removing them when they reach the distal end of the cell.
Proposition 4.1. Let ({(Qi(t), Pi(t)), i = 1, . . . , N})t≥0 be the number of parti-
cles in the axonal transport system with compartments of size δ, on and off transport,
14
respectively. Suppose the rate of production of particles from the source is rq0 = vρ0.
Then this Markov chain has the product-form stationary distribution
k1 (cid:19)
(Qi, Pi) ∼ P ois(q0) ⊗ P ois(cid:18) k2q0
where all {(Qi, Pi), i = 1, . . . , N} are mutually independent.
Proof. Since the production rate is rq0, the generator of the process (Q1, P1) is
Aq0 f (q, p) = [f (q + 1, p) − f (q, p)]rq0 + [f (q − 1, p) − f (q, p)]rq
+ [f (q − 1, p + 1) − f (q, p)]k2q + [f (q + 1, p − 1) − f (q, p)]k1p
If we use f (q, p) = Q1(t) and f (q, p) = P1(t), and take expectations, we get a system
of ODEs governing the change in E[Q1], E[P1] over time
dE[Q1](t)
dt
dE[P1](t)
dt
= rq0 − rE[Q1(t)] − k2E[Q1(t)] + k1E[P1(t)]
= k2E[Q1(t)] − k1E[P1(t)]
indicating that in equilibrium in the first section the mean numbers of on-transport
particles and off-transport particles are E[Q1] = q0, and E[P1] = E[Q1] k2
= q0k2/k1,
k1
respectively. Let πq0 (q, p) = πλQ (q)⊗ πλP (p) be a product of two independent Poisson
distributions with rates λQ = q0, and λP = q0k2/k1 respectively. To show that
πq0 (q, p) is a stationary distribution for the process (Q1, P1) we need to check that
∞
∞
Xq=0
Xp=0
Aq0 f (q, p)πq0 (q, p) = 0
for any choice of function f ∈ D(Aq0 ).
Aq0 f (q, p)e−(λQ+λP ) λq
Q
q!
λp
P
p!
∞
∞
Xq=0
Xp=0
= e−(λQ+λP )
+ [f (q − 1, p + 1) − f (q, p)]k2q + [f (q + 1, p − 1) − f (q, p)]k1p(cid:17)
λp
P
λq
Q
q!
∞
∞
Xq=0
Xp=0
p! (cid:16)[f (q + 1, p) − f (q, p)]rq0 + [f (q − 1, p) − f (q, p)]rq
λq+1
Q
(q + 1)!
λp
P
p!
λq
Q
q!
λp
P
p!
rq
r(q + 1) −
λq
Q
q!
λq
Q
q!
λp
P
p!
λp
P
p!
k2q
k1p
k2(q + 1) −
k1(p + 1) −
λQ
q + 1
p
λP
k2(q + 1)
λq−1
Q
(q − 1)!
λp
P
p!
rq0 −
rq0 +
λp
P
p!
λp−1
P
(p − 1)!
λp+1
P
(p + 1)!
λq
Q
q!
λq+1
Q
(q + 1)!
λq−1
Q
(q − 1)!
rq0 − rq0 +
λP
p + 1
q
λQ
+
+
λp
P
p! (cid:16) q
λQ
=
λq
Q
q!
=
λq
Q
q!
λQ
q + 1
r(q + 1) − rq +
k1(p + 1) − k1p(cid:17)
− k2q +
rq0 − rq0 + q0r − rq(cid:17) = 0
λp
P
p! (cid:16) q
q0
In the two sums the factor multiplying f (q, p) for any (q, p) ∈ N × N comes only from
terms involving {q − 1, q, q + 1} and {p − 1, p, p + 1} and equals e−(λQ+λP ) times
15
since λQ = q0 and λP /λQ = k2/k1.
Thus, in equilibrium the input rate for (Q2, P2), which is rQ1, has a Poisson
distribution with mean rq0, and is independent of P1. Let πq0 (q1)⊗ πλQ (q2)⊗ πλP (p2)
be a product of three Poisson distributions with rates q0, λQ = q0, and λP = q0k2/k1
respectively. To show that this is a stationary distribution for the process (Q1, Q2, P2)
we need to check that
∞
∞
∞
Xq1=0
Xq2=0
Xp2=0
Aq1 f (q2, p2)πq0 (q1)πq1 (q2, p2) = 0
for any choice of function f ∈ D(Aq1 ). For each fixed value q1, according to our
previous calculation the inner two sums give 0, so the whole sum is 0.
Thus, in equilibrium, (Q2, P2) have the distribution πλQ (q) ⊗ πλP (p) with λQ =
It follows by induc-
q0, λP = q0k2/k1 as well, and are independent from (Q1, P1).
tion that the stationary distributions for {(Qi, Pi)} are independent and identically
distributed as πλQ (q) ⊗ πλP (p), with λQ = q0, λP = q0k2/k1. This is also an exam-
ple of a clustering process satisfying the detailed balance conditions with linear rates
discussed in Sec. 8.2 of [25].
)q0 = (1 + k2
k1
We point out that the mean number of particles both on and off transport is
(1 + k2
)ρ0δ, scaling with the size of a compartment. To obtain the mean
k1
number of particles per unit length we add particles in ≈ 1/δ compartments, and the
mean number of particles per unit length is (1 + k2
)ρ0 independent of the choice of
k1
compartment size.
One immediate consequence is the analogous result for the number of particles
in the stochastic non-compartmental model at any location along the axon. Namely,
suppose the particles move according to the piecewise-linear Markov process (X, ξ)
with a Poisson rate ρ0 influx of new particles at location 0. Then, at any location
0 < x < L along the axon, the numbers of particles (Q(x,x+dx), P(x,x+dx)) on and
off transport, respectively, have the stationary distribution Pois(ρ0dx)⊗ Pois( k2
ρ0dx)
where for any x1, . . . xk ∈ (0, L), {Qxi}1≤i≤k and {Pxi}1≤i≤k are mutually indepen-
dent. We note that this result would not have been obvious to notice without going
through the compartmental model first, yet its consequences for prediction and anal-
ysis of the long term stochastic behavior of the system are quite powerful.
k1
16
4.2. Homogeneity of the axons at equilibrium. Recall that δ = 10nm,
roughly the step size of motor proteins, and that axons can be up to one meter in
length. Thus we are interested in phenomena on all the length scales 10νδ, where
ν = 0, 1, 2, . . . 8. Let ∆ = 10νδ; we want to determine how similar different segments
of the axon of size ∆ are. Let Q∆ and P∆ denote the numbers of on-track and off-track
particles in a segment of length ∆.
In equilibrium, Q∆ and P∆ are both sums of 10ν independent Poisson random
variables with parameters λQ = q0 and λP = k2
q0, respectively. Therefore the distri-
k1
butions of Q∆ and P∆ are Poisson with parameters 10νλQ and 10νλP , respectively.
The mean and the variance of the number of particles in the segment of length ∆ is
10ν(λQ + λP ). To see how homogeneous different slices of length ∆ are, we consider
the coefficient of variation, c∆, which is the standard deviation divided by the mean.
c∆ =
1
=
p(λQ + λP )10ν
p(1 + k2/k1)q010ν
1
.
As indicated in Section 2, q0 is in the range 0.25 to 2.5 in different axons. For illus-
trative purposes here, we will assume q0 = 1. Since k2/k1 = 3, we see that the scale-
dependent coefficient of variation c∆ = 1/(2√108∆). Therefore at the ten nanometer
scale the coefficient of variation is simply 1/2. At the micron scale c∆ = 1/20 and at
the millimeter scale c∆ = 0.5 × 10−5/2. The cutoff between "high variance" and "low
variance" distributions is usually considered to be when the coefficient of variation is
near 1, so by this standard the axon is extremely homogeneous in its length at large
scales.
4.3. Balance between efficient transport and targeted delivery. The pre-
ceding characterization of the transport apparatus enables us to address questions
concerning whole cell function. One core issue is that intracellular transport must
simultaneously accommodate two functional demands: some material, such as the
enzymes used to construct neurotransmitters, must be transported from the soma to
the axon terminal in a timely manner; whereas other cargo, such as sodium channels,
need to be delivered to unspecified locations as needs arise throughout the length of
the cell. The tradeoff between these two goals is clear. If a typical vesicle spends the
vast majority of its time in transport mode, the mean velocity will be close to the
mean on-transport velocity, but any needs that arise in the central part of the cell
will be neglected. On the other hand, if a typical vesicle spends too much time off
transport, presumably available for use if needed locally, then it will take substantially
longer to traverse the entire cell.
Recently Bressloff and Newby [31, 32] modeled particles that are created near
the nucleus that then undergo intermittent search (being in search mode while off
transport and not in search mode when on transport) for a target hidden somewhere
along the axon. Their model is a non-compartmental individual particle model with
the additional feature that vesicles can move backwards as well as forwards. They
compute the probability that the particle is successful and conditioning on success,
the mean first hitting time. With our system-wide model we can accommodate the
observation the if a given vesicle misses the target, another vesicle with similar cargo
will pass by before too long. We will assess this hitting time under two assumptions
about the density of relevant material. Our standing assumption q0 ∼ 1 is appropriate
for types of cargo that are found densely throughout the cell. In this setting, the wait
time is essentially just the time it takes for one of several nearby vesicles to unbind
17
from transport in the target region. A more interesting case is a setting where the
needed cargo is sparsely distributed, say q0 ∼ 10−3. In this setting, if the first cargo
to reach the target region fails to unbind, there will be significant time before the
next arrival. As we will see, we can still assess the trade-off intrinsic between risking
a target miss and diminishing the time of the next arrival.
To make the discussion concrete we define a target region Rn = {i∗ +1, . . . , i∗ +n}
where i∗ ∈ 1, . . . , N − n. At time zero, we take the system {Qi(0), Pi(0)} to be drawn
from the stationary distribution described by Proposition 4.1 conditioned on the event
that for all i ∈ Rn, Qi(0) = Pi(0) = 0. We introduce the hitting time Hn := inf{t >
0 : Pi∈Rn
Pi(t) > 0}, which marks the first time a particle is off transport while in
Rn. Since computing the mean of Hn is analytically intractable, we introduce another
hitting time H ′
n, stochastically dominating Hn, that nevertheless reflects the essential
tradeoff between maximizing mean velocity and making detachment from transport
likely in the target region.
Let I∗ := {i ∈ {1, . . . , i∗} Qi(0) + Pi(0) > 0} be the set of all non-empty sections
of the cell at time 0. Among the particle in these sections some will be "successful"
in that they will detach from transport in the target region, while others will be
unsuccessful. Labeling each particle in these terms, we decompose I0 into locations
with successful particles I s
∗ . We are interested in the time
H ′
n at which the rightmost successful particle, that is, a particle starting from position
im = max{I s
∗}, detaches while in Rn. If S = 0 then we define im = max{∅} := 0.
For the position of this particle we use the notation (X δ
t )t≥0 from Section 3, where
X δ
0 = δim, X δ
t ∈ {δim, . . . , δN}, together with the indicator (ξδ
t ∈ {0, 1} of
whether the particle is on or off transport, respectively. Let
∗ and unsuccessful particles I u
t )t≥0, ξδ
H ′
n := inf{t > 0 : (X δ
t , ξδ
t ) = (δ(i∗ + 1), 1)}.
∗ , H i
n = inf{t > 0 : X δ
∗ where for each i ∈ I s
n : i ∈ I s
Note that H ′
t ∈ δRn}, since particles cannot skip sections, and
always enter a section on transport. If we were to analogously define a sequence of
times {H i
n}i∈I s
0 = δi
and ξδ
0 ∈ {0, 1}, then the exact first hitting time of the target region will satisfy
Hn = min{H i
Since the exact distribution of H i
n is complicated, E[Hn] is analytically intractable,
and instead we focus on finding a simple expression for E[H ′
n]. Our point is that, in
the sparse material limit (q0 small), the rightmost particle becomes increasingly likely
to be the first successful particle to detach in the target region, and Hn approaches
H ′
n.
∗}. Hence, clearly Hn ≤ H im
n = inf{t > 0 : X δ
t ∈ δRn} with X δ
n ≡ H ′
n.
The computation of E[H ′
n] requires computing the time it takes a particle to travel
a certain axonal distance, given by the following.
Lemma 4.2. Let the initial position of the particle be (X δ
0) = (0, 1) and let
L∗ ∈ {1, . . . , L} be given where L is the total length of the axon. Then, the time
TL∗ = inf{t > 0 : X δ
t = L∗}, satisfies
0 , ξδ
E[TL∗] =
L∗(k2 + k1)
vk1
.
Proof. Since (X δ
t , ξδ
t )t≥0 is a Markov process with generator
Aδf (x, ξ) = rξ(cid:2)f (x + δ, ξ) − f (x, ξ)(cid:3)
+ k2ξ(cid:2)f (x, ξ − 1) − f (x, ξ)(cid:3) + k1(1 − ξ)(cid:2)f (x, ξ + 1) − f (x, ξ)(cid:3)
18
it follows that M 1
s )ds are
both martingales. Since both M 1 and M 2 have bounded increments and E[TL∗] < ∞,
the optional stopping theorem implies that
0 k1(1 − ξδ
t := X δ
t := ξδ
0 ξδ
s ds and M 2
t − rδR t
TL∗i = L∗ − vEhZ TL∗
0
0 k2ξδ
s ds −R t
t +R t
s dsi ⇒ EhZ TL∗
ξδ
0
ξδ
0 = E(cid:2)M 1
0(cid:3) = EhM 1
and
s dsi = L∗/v
1 = E(cid:2)M 2
0(cid:3) = EhM 2
TL∗i = 1 − k1E[TL∗] + (k2 + k1)EhZ TL∗
0
ξδ
s dsi
which implies
E[TL∗] =
k2 + k1
k1
EhZ TL∗
0
ξδ
s dsi.
and our claim follows.
We also note that the same computation holds for the mean time a particle in
the stochastic non-compartmental model (Xt, ξt)t≥0 takes to reach a distance L∗, as
the two martingales used in the proof depend only on v and not on δ.
We next compute the hitting time H ′
n of a particle started at location im ∈ I∗.
Lemma 4.3. Let the system {(Qi(0), Pi(0)), i = 1, . . . , N} have the stationary
distribution given by Proposition 4.1 conditional on Qi(0) = Pi(0) = 0, for all i ∈ Rn.
Then
k2 + r(cid:19)(cid:21)
n] = (1 − e−λn)(cid:20)
k2 + r(cid:19)n(cid:18)1 +
k2pn (cid:20)1 −(cid:18) r
k1(k2 + k1)(cid:21) +
rq0pn
E[H ′
nk2
k2
+
1
1
r
k2+r )n and λn = k2+k1
where pn = 1 − (
Proof. The proof of Proposition 4.1 shows that conditioning on the values of
Pi(0), i ∈ Rn does not affect the law of {(Qi(t), Pi(t)), i ≤ i∗}, hence for any t ≥ 0,
they form two mutually independent sequences of Pois(q0) and Pois( k2q0
) random
k1
variables.
q0i∗pn.
k1
k1
probability it first detaches at location i∗ +i. Since p(i) = (
detaching there. This probability can be written pn := Pn
a particle gets off transport in the i-th compartment, pn =Pn
Let S denote the number of successful particles between sites 0 and i∗ at time
zero. The total of number of particles at time 0 at these sites that are either on
or off transport is distributed as a Pois( k2+k1
q0i∗) variable. We next compute the
probability pn that any particle once it reached the target region is "successful" in
i=1 p(i) where p(i) is the
k2+r )i−1 k2
r
k2+r is the chance
k2+r )n.
The probability that a particle is successful does not depend on which location it
was at time 0, and whether it was on or off transport at that time. Hence, S is is
distributed as a Pois( k2+k1
q0i∗pn) variable, and conditioned on the value S, the set
of locations of the particles at time 0 is distributed as a set of S draws from the
uniform distribution on {1, . . . , i∗}. Since im is the maximum of S samples from a
Uniform{1, . . . , i∗} distribution we have E[imS] = i∗ − 1
x=1 xS ≈ i∗
e,n is the time it takes a
successful particle initially at location im to enter the region Rn, and Ho,n is the time
it takes any successful particle after it enters the region to get off transport. Thus,
E[H ′
o,n, where we let H ′
∗ Pi∗−1
We now decompose H ′
i=1 p(i) = 1 − (
e,n + H ′
n = H ′
S
S+1 .
k1
iS
r
n] = E(cid:2)H ′
e,n(cid:3) + E(cid:2)H ′
o,n(cid:3).
If a successful particle starts at location X δ
ξδ
0 = 1, then the time it takes to enter the region Rn is by Lemma 4.2
0 = im ∈ {0, . . . , i∗} on transport
19
e,n(X δ
0 , ξδ
E(cid:2)H ′
0) = (im, 1)(cid:3) =
(i∗ − im)δ(k2 + k1)
vk1
If it starts at location X δ
0 = 1 then it takes an
additional Exponential time with mean 1/k1 for it to get back on transport at the
, and
0 = im ∈ {1, . . . , i∗} off transport ξδ
0 , ξδ
since we assume new particles always enter the system on transport
0 , ξδ
e,n(X δ
same location, so E(cid:2)H ′
E(cid:2)H ′
S+1 , P{im > 0} = P{S > 0}, we get that
0) = (im, 0)(cid:3) = E(cid:2)H ′
(i∗ − im)δ(k2 + k1)
e,nim(cid:3) =
e,n(X δ
1
k1
vk1
+
S
1im>0.
0) = (im, 1)(cid:3) + 1
k1
Since E[imS] = i∗
E(cid:2)H ′
e,n(cid:3) = E(cid:20)
with S ∼ Pois(λn), λn = k2+k1
we get
k1
1
S + 1(cid:21) i∗(k2 + k1)
rk1
+
k2
k1(k2 + k1)
P{S > 0}
q0i∗pn. Since E[S] = 1−e−λn
λn
, P{S > 0} = 1 − e−λn ,
+
k2
rq0pn
k1(k2 + k1)(cid:19)
e,n(cid:3) = (1 − e−λn )(cid:18) 1
E(cid:2)H ′
To calculate E(cid:2)H ′
o,n(cid:3) note that if a particle first gets off transport in the i-th
compartment then the time of its travel until this point is a sum of i independent
and identically distributed exponential random variables with parameter k2 + r. The
probability a successful particle first gets off transport in the i-th compartment is
p(i)
. Hence, the time a successful particle takes to get off transport once it enters the
pn
region R has the mean
E(cid:2)H ′
o,n(cid:3) =
p(i)
pn
n
Xi=1
i
k2 + r
=
and our claim follows.
1
k2pn (cid:20)1 −(cid:18) r
k2 + r(cid:19)n(cid:18)1 +
nk2
k2 + r(cid:19)(cid:21) .
To complete our analysis, we wish to characterize this result in terms of length
along the axon and independent of the stepping size δ. To this end, we fix a length ℓ
and for a given compartment size ℓ, we let n = ⌈ℓ/δ⌉, and for the start of the region
we let i∗ = ⌈ℓ∗/δ⌉. As such, as δ → 0 the limiting region becomes Rℓ = (ℓ∗, ℓ∗ + ℓ).
Then, under the assumption that q0/δ → ρ0 we have
pn → pℓ = 1 − e−ℓk2/v, λn → λℓ =
k2 + k1
ρ0ℓ∗pℓ
k1
and E[H ′
n] → E[H ′] where
E[H ′] =(cid:0)1 − e−λℓ(cid:1)(cid:20) 1
vρ0pℓ
+
k2
k1(k2 + k1)(cid:21) +
1
k2pℓ (cid:20)1 − e− ℓk2
v (1 +
ℓk2
v
)(cid:21) .
(4.1)
It remains to interpret this result with respect to the parameter choices we have
made. First, we must set a value for the size ℓ of the target region. For this purpose
20
we note that the typical size of a Node of Ranvier -- a gap in the myelin sheath of a
myelinated axon where sodium channels are concentrated in the cell membrane -- is
approximately one micron.1
As seen in the preceding proof there are three contributions to E[H ′], and we next
analyze them in terms of their dominance for the overall value. We begin with the
contribution from the last term, which is the time it takes for a successful particle to
detach once it has reached the target region. By our choice of ℓ, the recurring ratio
ℓ/v is one. Along with the earlier assumption that k2 = 1, the entire term simplifies
to (e− 2)/(e− 1) = 0.4 s. The multiplicative factor 1− e−λℓ preceding the first term of
(4.1) results from a boundary effect: when ℓ∗ is very close to zero, it is very unlikely
there are any particles already in the system between the soma and the target region.
When the target region is in the middle of the axon, this contribution from particles
in that section of the axon at time 0 is significant.
When q0 ∼ 1 as assumed earlier, then ρ0 = q0/δ ∼ 108. Then if ℓ∗ > 10−8, which
corresponds to the target region being just one motor step down the length of the
axon, we have ρ0ℓ∗ ∼ 1 and λℓ > 2.3 so 1 − e−λℓ > 0.9. Looking at the first term
inside the parenthesis, we note that vρ0 ∼ 102, while pℓ = 1 − exp(−ℓk2/v) ≈ 0.6 is
the probability that any given vesicle will be successful in detaching from transport in
the target region, so 1/(vρ0pℓ) ∼ 10−2. Meanwhile, the term k2/(k1(k2 + k1)), which
is the expected time to bind to transport if a successful particle happens to be off
transport at time 0, is 1/4 for the assumed values of k2 and k1. Therefore under the
q0 ∼ 1 assumption, both this and the final term contribute significantly to the hitting
time.
However, in the sparse material regime, say q0 ∼ 10−3, we have then ρ0 ∼ 105,
and the factor 1 − e−λℓ > 0.9 when ℓ∗ > 10−5. That is, if the target region is at
least 1/100, or at least 1000 segments, down the length of a 10−3m axon. However,
now the rate limiting factor is the wait time for the first successful vesicle to arrive
in the target region, which is captured by the term 1/vρ0p. The product vρ0 ≈ 0.1
measures the average rate at which new particles should arrive and together with the
probability of success pℓ ≈ 0.6, we now have 1/vρ0pℓ ≈ 16s. So, in the sparse material
regime the first summand in E[H ′] giving the mean time for arrival of the particles to
the target region dominates.
It is interesting to consider what happens to the arrival rate term under per-
turbations of the various parameters. In particular, we draw the reader's attention
to changes in v. When viewing intermittent search in terms of a single particle, the
probability of finding the target is strictly decreasing in v. Higher velocity seems to be
the enemy of finding the target. Indeed, from a system point of view, this implies that
the system will require more trials before a successful particle arrives in the target
region. What the equation (4.1) gives us is the ability assess how much more quickly
the trials will happen. In fact, the function v(1 − exp(−ℓk2/v)) is increasing in v.
Therefore the entire expected wait time E[H ′] is actually decreasing in v. That is to
say, while the particles are less likely to succeed, they will be arriving more rapidly
enough to counterbalance the lost time. We believe that this kind of quantitative
analysis will prove fruitful in future study when coupled with more details of the
1We do not wish to claim this is a complete model for deposition of sodium channels in Nodes of
Ranvier, as the particulars of the biology -- which may include factors like local signals that encourage
motors to detach from microtubules near the Nodes -- are not fully understood. We merely wish to
use the size of the Nodes to fix our intuition about the size of a target to other important length
scales such as that of microtubules which are also a micron in length.
21
biology of deposition of materials in the cell membrane.
4.4. Approaching Equilibrium. We have seen above that the axon is very
homogeneous at stochastic equilibrium on a space scale down to micrometers. One
of the beautiful properties of transport with reversible binding is that if it is locally
out of equilibrium, the on-off dynamics can return the system to equilibrium on a
much faster time scale than waiting for new material to arrive from the nucleus.
Furthermore, as is discussed in [45], one of the key goals of biophysics investigation is
the discovery of behaviors for which biochemical regulation is not necessary. This is
of fundamental importance to the biological function of the system because it means
that the axon will automatically "repair" itself without central control of the repair
process.
How good is this mechanism? If a segment of the axon is far away from equilib-
rium, how long does it take to get back close to equilibrium? To investigate this ques-
tion, we imagine that the axon is at stochastic equilibrium except for some segment
R, where the total number of particles on and off transport, Qi(0)+ Pi(0) = 0,∀i ∈ R,
is zero at time 0. Let a and ε be given small numbers and suppose that λ∞ is the
mean vector for (Qi, Pi) at stochasttic equilibrium. We want to compute (an upper
bound for) the time t∗ so that
P{(Qi(t∗), Pi(t∗)) − λ∞ ≥ aλ∞,∀i ∈ R} ≤ ε,
that is, the probability that (Qi, Pi),∀i ∈ R is significantly different from λ∞ is very
small. In the applications below we will choose a = 0.1 and ε = 0.05, and we will
see that a 10 micron segment can recover in about 10 seconds, while a 1 millimeter
segment will take about 1000 seconds or 15 minutes to recover. We study this question
first for a single location R = {i}, and then use the estimates derived to scale the
results to segments of any length.
Proposition 4.4. Let (Qi(0), Pi(0)) = (0, 0) and let the constants a > 0 and
) is the equilibrium
ε ∈ (0, 1) satisfy the relationship a2ελ∞ > 1 where λ∞ = q0(1, k2
vector of (Qi, Pi). Then there exists t∗ > 0 such that ∀t ≥ t∗,
P{(Qi(t), Pi(t)) − λ∞ ≥ aλ∞} ≤ ε.
√pλ∞
a√ελ∞−1(cid:17) is sufficient, where
2(cid:16)k2 + k1 + r −p(k2 + k1 + r)2 − 4k1r(cid:17) .
In fact, the choice t∗ = α−1 ln(cid:16)
α :=
k1
1
guarantees that there are "enough" particles.
We note that given a particular choice of a and ε, the condition a2ελ∞ > 1
Proof. Note that for any given β ∈ (0, 1), we may choose t∗ > 0 such that for all
t ≥ t∗, the vector of means λ(t) := E[(Qi(t), Pi(t))] satisfies λ(t) − λ∞ ≤ aβλ∞.
Then
P{(Qi(t), Pi(t)) − λ∞ ≥ aλ∞} ≤ P{(Qi(t), Pi(t)) − λ(t) + λ(t) − λ∞ ≥ aλ∞}
≤ P{(Qi(t), Pi(t)) − λ(t) ≥ a(1 − β)λ∞}
Applying Chebyshev's Inequality, and observing that the variance of a Poisson random
variable is equal to its mean, we conclude that
P{(Qi(t), Pi(t)) − λ∞ ≥ aλ∞} ≤
Var[(Qi(t), Pi(t))]
a2(1 − β)2λ∞2 =
λ(t)
a2(1 − β)2λ∞2
22
Since the initial condition for both Pi and Qi are less than their respective equilibrium
values, each are monotonically increasing functions and the above reduces to
P{(Qi(t), Pi(t)) − λ∞ ≥ aλ∞} ≤
1
a2(1 − β)2λ∞
for all t > t∗. In order to satisfy the requirement that the right hand side must be less
than ε, we solve for β and find β = 1 − (apελ∞)−1 provided that apελ∞ > 1.
It remains to study the convergence of the mean and the appropriate choice of
t∗. The dynamics of the mean vector λ(t) are given by the ODE
(4.2)
λ(t) = −A1λ(t) + q0re1
where e1 is the unit vector (1, 0) and A1 =(cid:18) k2 + r −k1
−k2
k1 (cid:19) .
d
dt
The solution to (4.2) is λ(t) = λ∞ + e−A1t(λ(0) − λ∞) where λ∞ = q0rA−1
1 e1 =
). This yields the estimate
q0(1, k2
k1
λ(t) − λ∞ ≤(cid:12)(cid:12)e−A1tλ∞(cid:12)(cid:12) ≤ e−αtλ∞
where α is the smaller of the eigenvalues of A1. Noting that α > 0, t∗ may be chosen
so that e−αt∗ ≤ aβ, i.e. t∗ = α−1 ln(1/(aβ)) is an upper bound for the time to be
close to equilibrium with high probability.
In order to calculate return to equilibrium at various scales, we now suppose
that the whole axon is in statistical equilibrium except for a segment R of length
∆ = δ10ν in which we will assume that there are no particles either on or off the
track. Proposition 4.4 covered the case ν = 0. We are interested in ν = 1, . . . , 8. We
imagine that the axon is broken up into 108−ν segments of length ∆. In this rescaled
system, the unbinding and binding rates per particle, k2 and k1 remain the same, as
well as the mean on-transport velocity v. In order to retain this mean velocity, the
rate of lateral stepping must be decreased to r = r10−ν.
The ODE for the mean vector of the rescaled system is given by
d
dt
λ(t) = − A1 λ(t) + q0re1
(4.3)
with A1 =(cid:18) k2 + r −k1
k1 (cid:19) .
−k2
We note that the last term in (4.3) contains an r rather than an r. This is because
the input rate is unchanged while the exit rate is diminished. The resulting equilibrium
value is therefore rescaled as well, λ∞ = q0r A−1
1 e1 = q0
components of this vector are of order 10ν, as expected.
k2/k1(cid:1) = q010ν(cid:0) 1
r(cid:0) 1
Using the parameters discussed in Section 2: k2 = 1, k1 = 1
3 , v = 10−6m/s, r =
102s−1. This implies r = 102−νs−1. We choose the thresholds to be a = 0.1 and
ε = 0.05, so the constraint that a2ελ∞ > 1 requires that λ∞ > 2 × 103. Since
0.25 ≤ q0 ≤ 2.5, this is indeed the case if ν > 3, i.e. if the segment has length greater
than 10 microns.
It follows from Proposition 4.4 that the time to equilibrium, t∗, is proportional
k2/k1(cid:1). Both
r
to α−1 where α satisfies:
1
α =
=
2(cid:16)k2 + k1 + r −p(k2 + k1 + r)2 − 4k1r(cid:17)
k2 + k1 + r +p(k2 + k1 + r)2 − 4k1r
2k1r
.
23
For the given parameter values (with λ∞ > 5 × 103 in particular), the constant
of proportionality ln(cid:16)pελ∞/(apελ∞ − 1)(cid:17) is contained in the interval (2.3, 2.5)
and does not have a impact on how the relaxation time scales with ν.
In terms
of analyzing α, we note that k1r is small compared to k2. To leading order, α ∼
(k1r)/(k2 + k1 + r) ∼ 102−νs−1. It follows that t∗ ∼ 10ν−2s. Thus, for a 10 micron
segment (ν = 3) the time to recover is about 10 seconds and for a 1 millimeter seg-
ment (ν = 5) the time to recover is 1000 seconds or 15 minutes. The time to recover
depends, of course, on the parameter ε that represents what we mean by "close." We
also note that one can compute various measures of time to recover using the PDE
models discussed in Section 3.
5. Discussion. In this paper, we created a spatial Markov chain model for study-
ing various aspects of fast axonal transport. Previous models that use PDEs treat
the velocity of transport as constant when particles are attached to the fast transport
system. Since it is known that transport along the microtubules is itself stochastic,
it is important to have a fully stochastic model. Our model allows us to unify and
extend previous work. In Section 3.2 we show that from the particle perspective as
the compartment size δ tends to zero, our model converges in distribution on the
Skorokhod space of c`adl`ag functions to the piecewise-deterministic model analyzed
by Brooks [5]. Namely, we show that the paths of particles in our model converge to
those of particles in a stochastic non-compartmental model. The argument proceeds
by an explicit computation of the finite-dimensional distributions and a tightness ar-
gument. In Proposition 3.4 we give a rigorous probabilistic proof of why the paths of
particles follow "approximate traveling waves" described by other authors [36, 5, 14].
This proof is based on stochastic averaging arguments which show that a functional
central limit theorem holds on the space of continuous functions for the paths of
particles as the compartment size decreases. The diffusion of particles around their
mean position can consequently be approximately described jointly for all time by a
Brownian motion with the appropriate diffusion coefficient.
In Section 2, we show how to use existing experimental data to indentify (ranges
for) all the parameters of our model.
In light of this, we can use the model to
investigate several important biological questions. These are based on describing the
spatial distribution of multiple particles in our model. In Section 4.1 we derive the
stationary distribution for the number of particles in different compartments on and
off transport along the axon. This gives an explicit description of the stochasticity of
the system that is present even after a long time. In Section 4.2 we derive estimates
for how homogeneous the axon is on different spatial scales. In Section 4.3 we study a
question introduced by Bressloff, by providing a stochastic quantity which describes
the balance the system needs to achieve between rapid transport that brings new
material quickly and efficient local search that improves time of delivery to a target.
Finally, in Section 4.4, we use the model to calculate the length of time that it would
take for axonal segments of different lengths to recover to near stochastic equilibrium
after they have been depleted of vesicles.
In our stochastic compartmental model all event wait times are assumed to be
exponential random variables, but this is certainly a simplification. As an example,
the stepping process of kinesin is a well-studied though still not completely understood
phenomenon. Much work has focused on assessing the dependence of the mean rate of
translocation on both the load and the local concentration of ATP [44] [38]. Implicit
in this analysis is the assumption of exponential wait times with state dependent rate
24
parameters. However, when fitting to data and matching dispersion information the
authors in [13] found it necessary to generalize the wait time distribution. This was
followed by more detailed models for which it was shown that load carrying could
in fact regularize the stepping times of kinesin motors [37] [9]. Generalizing waiting
times would significantly affect our results. Since the particle position process is no
longer Markov, we no longer have the direct connection to the previous results stated
in Section 3.1.-3.3., nor can we use the stationary distribution employed in Section
4.1 and used for addressing the biologicals questions in Sections 4.2.-4.4. In light of
the known need for generalized wait times in the stepping process, it seems likely that
detailed observation of the rebinding process will call for new mathematical models as
well. Recall that when a vesicle unbinds from a microtubule it is unclear whether it
typically rebinds to the same microtubule or if it explores the region significantly via
diffusion before finding a different microtubule to bind to. In the latter case, a more
appropriate model for rebinding time would be to solve some kind of first passage
time problem and use that distribution for the rebinding wait.
An important aspect of the biology of axonal transport is not included in the
model presented here, namely the local deposition and eventual degradation of trans-
ported materials. For example, sodium channels and sodium pumps are synthesized
at the soma, transported down the axon and deposited in the axonal membrane, ether
uniformly as in an unmyelinated axon or at the nodes of Ranvier in a myelinated axon.
Channels and pumps are proteins with half-lives on the order of days to weeks. The
present model can be extended to include a deposition compartment at each location,
and, clearly, the processes of deposition and subsequent degradation will cause the
mean number of particles both on and off transport to be monotone decreasing as
one moves down the axon. How inhomogeneous this makes the axon will depend on
the details of deposition and degradation rates. Our preliminary calculations indi-
cate that long axons, such as the meter-long axons in human sciatic nerve, would be
quite inhomogeneous. This is an important biological issue because it is controversial
whether the machinery for protein synthesis (i.e. ribosomes) exist in axons [42]. We
have also not included retrograde transport or the fact that some axons may have
location-dependent unbinding rates [10]. All of these issues will be the subject of
future work.
Acknowledgments. The authors are grateful to Professors H. Frederik Nijhout
and Vann Bennett of Duke University and Professor Anthony Brown of The Ohio
State University for helpful discussions. This work was supported by NSF grants
DMS-061670 and EF-1038593, and an NSERC Discovery Grant.
REFERENCES
[1] Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2008) The Molecular Biology of
the Cell, Garland Science, Taylor & Francis Group, New York.
[2] Allen RD, Weiss DG, Hayden JH, Brown DT, Fijiwaki H, Simpson M (1985). Gliding movement
and bidirectional transport along single native microtubules from squid axoplasm: evidence
for an active role for microtubules in cytoplasmic transport. J. Cell Biol 100, 1736-1752.
[3] Blum JJ, Reed MC (1985) A Model for Fast Axonal Transport. Cell Motility 5, 507-527.
[4] Bressloff PC, (2006) Stochastic model of protein receptor trafficking prior to synaptogenesis
Physical Review E 74 031910.
[5] Brooks EA (1999) Probabilistic methods for a linear reaction-hyperbolic system with constant
coefficients. Ann. Appl. Prob., 9(3), 719-731.
[6] Brown A (2000) Slow axonal transport: stop and go traffic in the axon. Nat Rev Cell Mol Biol
1, 153-156.
25
[7] Carter NJ, Cross RA (2005) Mechanics of the kinesin step. Nature 435, 308-312.
[8] Cox DR (1962) Renewal Theory. Methuen & C0. Ltd., London, Section 4.5.
[9] DeVille RL and Vanden-Eijnden E (2008) Regular Gaits and Optimal Velocities for Motor
Proteins Biophysics Journal 95 6: 2681-2691.
[10] Dixit R, Ross JL, Goldman YE, Holzbaur ELF (2008) Differential regulation of dynein and
kinesin otor proteins by tau. Science 319, 1086-1089.
[11] Ethier S, Kurtz TG (1986) Markov Processes: Characterization and Convergence. John Wiley
& Sons Inc., New Jersey
[12] Finer JT, Simmons RM, Spudich JA (1994) Single myosin molecule mechanics: piconewton
forces and nanometre steps. Nature 368, 113-119.
[13] Fisher ME, Kolomeisky (2001) Simple mechanochemistry describes the dynamics of kinesin
molecules. PNAS 98 14: 7748-7753.
[14] Friedman A, Craciun G (2005) A model of intracellular transport of particles in an axon. SIAM
J Appl Math 38, 741-758.
[15] Friedman A, Craciun G (2006) Approximate traveling waves in linear reaction-hyperbolic
equations. SIAM J Appl Math 38, 741-758.
[16] Friedman A, Hu B (2007) Uniform convergence for approximate traveling waves in reaction-
hyperbolic systems. Indiana Math J 56, 2133-2158.
[17] Friedman A, Hu B (2007) Uniform convergence for approximate traveling waves in reaction-
diffusion-hyperbolic systems. Arch Rat Mech Anal 186, 251-274.
[18] Gennerich A, Carter AP, Reck-Peterson, Vale RD (2007) Force-induced bidirectional stepping
of cytoplasmic dynein. Cell 131, 952-965.
[19] Griego RJ, Hersh R (1969) "Random evolutions, Markov chains, and systems of partial differ-
ential equations." Proc. Nat. Acad. Sci. USA 62, 305-308.
[20] Gross GW, Weiss, DG (1982) "Theoretical considerations on rapid transport in low viscosity
axonal regions." Axoplasmic Transport (Ed. D.G. Weiss), Springer-Verlag, Berlin.
[21] Jastrow H Electron Microscopic Atlas of Cells, Tissues and Organs on the Internet. Schwann
cells section.
http://www.uni-mainz.de/FB/Medizin/Anatomie/workshop/EM/EMSchwannE.html
[22] Kallenberg O (2002) "Foundations of modern probability", 2nd ed. Springer-Verlag, New York.
[23] Khas'minskii, RZ (1966) "On stochastic processes defined by differential equations with a small
parameter." Theory Probab. Appl. , 11, 211-228.
[24] Khas'minskii RZ (1966) "A limit theorem for the solutions of differential equations with random
right-hand sides." Theory Probab. Appl. , 11, 390-406.
[25] Kelly FR Reversibility and stochastic networks. John Wiley & Sons Ltd., Chichester, 1979.
Wiley Series in Probability and Mathematical Statistics.
[26] Kurtz TG (1992) Averaging for martingale problems and stochastic approximation. Lecture
Notes in Control and Inform. Sci. , 177, 186-209.
[27] Lasek RJ, Brady ST (1982). The structural hypothesis of axonal transport: two classes of
moving elements. Axoplasmic Transport (ed. G. Weiss), Springer-Verlag, Berlin, 1982, pp.
397-405.
[28] Lawler G (1995) Introduction to Stochastic Processes Chapman and Hall, New York.
[29] Miller R, Lasek RJ (1985) Crossbridges mediate anterograde and retrograde vesicle transport
along microtubules in squid axoplasm. J Cell Biol 101, 2181-2193.
[30] Moran D, Rowley JC III Visual Histology.com, Chapter 8, Nerves.
http://www.visualhistology.com/products/atlas/VHA Chpt8 Nerves.html
[31] Newby JM, Bressloff PC Directed Intermittent Search for Hidden Targets New Journal of
Physics 11 023033.
[32] Newby JM, Bressloff PC Quasi-steady State Reduction of Molecular Motor-Based Models of
Directed Intermittant Search Bullentin of Mathematical Biology 72 1840-1866.
[33] Ochs S (1972) Rate of fast axoplasmic transport in mammalian nerve fibers. J Physiol 227,
627-245.
[34] Reed MC, Blum JJ (1986) Theoretical analysis of radioactivity proles during fast axonal
transport: effects of deposition and turnover. Cell Motility and the Cytoskeleton, 6, 620627.
[35] Reed MC, Blum J (1994) Mathematical Questions in Axonal Transport. In Lectures in math-
ematics in the Life Sciences,24, Amer. Math. Soc., Providence.
[36] Reed MC, Venakides S, Blum JJ (1990) Approximate traveling waves in linear reaction-
hyperbolic equations. SIAM J. Appl. Math., 50, 167180.
[37] Schilstra MJ, Martin SR (2006) An elastically tethered viscous load imposes a regular gait on
the motion of myosin-V. Simulation of the effect of transient force relaxation on a stochastic
process. J.R. Soc. Interface 3 153-165.
[38] Schnitzer MJ, Visscher K, Block SM (2000) Force production by single kinesin motors. Nat.
26
Cell. Biol. 2 718-722.
[39] Stewart GH, Horwitz B, Gross GW (1982) A chromatographic model of axoplasmic transport.
In Axoplasmic Transport (ed. G. Weiss), Springer-Verlag, Berlin, 1982, 414-422.
[40] Svoboda K, Schmidt C, Schnapp B (1985) Direct observation of kinesin stepping by optical
trapping interferometry. Nature 365, 721-727.
[41] Takenaka T, Gotoh H (1984) Simulation of axoplasmic transport. J Theor Biol 107, 579-601.
[42] Twiss JL, Fainzilber M (2009) Ribosomes in axons - scrounging from the neighbors. Trends in
cell biology. 19, 236-243
[43] Vale RD, Reese TS, Sheetz MP (1985).
Identification of a novel force-generating protein,
kinesin, involved in microtubule-based motility. Cell 42, 39-50.
[44] Visscher K, Schnitzer MJ, Block SM (1999) Single kinesin molecules studied with a molecular
force clamp Nature 400, 184-189.
[45] Welte MA, Gross SP (2008) Molecular motors: a traffic cop within? HFSP Journal 2 4:
178-182.
[46] Willis WD, Grossman RG (1981) Medical Neurobiology: Neuroanatomical and Neurophysio-
logical Principles Basic to Clinical Neuroscience. Mosby, Inc.
|
1310.0218 | 1 | 1310 | 2013-10-01T10:02:24 | Synaptic metaplasticity underlies tetanic potentiation in Lymnaea: a novel paradigm | [
"physics.bio-ph",
"cond-mat.dis-nn",
"q-bio.NC"
] | We present a mathematical model which explains and interprets a novel form of short-term potentiation, which was found to be use-, but not time-dependent, in experiments done on Lymnaea neurons. The high degree of potentiation is explained using a model of synaptic metaplasticity, while the use-dependence (which is critically reliant on the presence of kinase in the experiment) is explained using a model of a stochastic and bistable biological switch. | physics.bio-ph | physics |
1
Synaptic metaplasticity underlies tetanic potentiation in
Lymnaea: a novel paradigm
Anita Mehta1,∗, Jean-Marc Luck2, Collin C. Luk3, Naweed I. Syed3
1 S. N. Bose National Centre for Basic Sciences, Calcutta, India
2 Institut de Physique Th´eorique, CEA Saclay and CNRS URA 2306, Gif-sur-Yvette,
France
3 Hotchkiss Brain Institute, Faculty of Medicine, University of Calgary, Alberta, Canada
∗ E-mail: [email protected]
Abstract
We present a mathematical model which explains and interprets a novel form of short-term potentiation,
which was found to be use-, but not time-dependent, in experiments done on Lymnaea neurons. The high
degree of potentiation is explained using a model of synaptic metaplasticity, while the use-dependence
(which is critically reliant on the presence of kinase in the experiment) is explained using a model of a
stochastic and bistable biological switch.
Introduction
All brain functions, ranging from simple reflexes to complex motor patterns, learning and memory, rely
upon synaptic transmission between neurons through specialized structures termed synapses. These
synaptic connections are, however, not static in nature; in fact, they exhibit a high degree of synaptic
plasticity, enabling a network to generate behaviorally relevant and functionally meaningful outputs.
These changes in synaptic strength can either be short- or long-term, and may form the basis for both
short- and long-term memory, respectively.
Synaptic plasticity is neither restricted to any select group of neurons nor to a particular species --
rather it is a universal trademark of all neurons that have been investigated to date. When a nervous
system is unable to exhibit modulatory changes associated with short- and long-term synaptic plasticity,
it is rendered dysfunctional. Therefore, defining the mechanisms underlying synaptic plasticity is not only
pivotal for our understanding of how the brain functions but also for managing the behavioral, learning,
memory and cognitive defects that are met in clinical practice. However, despite recent advances in
our understanding of various modes of neuronal communication, the cellular and molecular mechanisms
underlying synaptic plasticity remain poorly defined. Moreover, the data generated from experimental
studies has often been inadequate to garner mathematical modeling predictions that may aid future
research in this area, or to provide insights into the mechanisms of synaptic plasticity. This field could
however benefit from a paradigm shift where modeling approaches could be used to predict elements of
synaptic plasticity and to facilitate future research in the area of metaplasticity.
In a recent study [1], two of the authors of this paper observed a form of short-term potentiation
induced by tetanic stimulation, whose time-frame exceeded conventional forms of short-term potentiation.
While the induction of potentiation was similar to previous forms of short-term potentiation, the time
frame of the potentiated response was characteristic of long-term potentiation (∼ 5 hrs).
More specifically, using well-defined, excitatory cholinergic synapses between Lymnaea pre- and post-
synaptic neurons, specifically visceral dorsal 4 (VD4) and left pedal dorsal 1 (LPeD1-Excitatory), they
provided evidence for a novel form of short-term potentiation, which was use-, but not time-dependent.
They found that following a tetanic stimulation (∼ 10 Hz) in the presynaptic neuron with a minimum
of seven action potentials, the synapse became potentiated, whereby a subsequent action potential trig-
gered in the presynaptic neuron resulted in an enhanced postsynaptic potential (see Figure 1). Further,
if an inducing tetanic stimulation was activated but a subsequent action potential was not triggered, the
2
synapse was shown to remain potentiated for as long as 5 hours. However, once this action potential was
triggered, the authors found that the synaptic strength rapidly returned to baseline levels. It was also
shown that this form of synaptic plasticity relied on the presynaptic neuron, and required pre- but not
postsynaptic Ca2+/calmodulin dependent kinase II (CaMKII) activity. Hence, this form of potentiation
shares induction and de-potentiation characteristics similar to other forms of short-term potentiation, but
exhibits a time-frame analogous to that of long-term potentiation. The model of metaplastic synapses [2]
reviewed below allows us to reproduce these long timescales, and forms the basis of our explanation of
the experimental results reported in [1].
In the model of metaplastic synapses, incoming signals are stored as memories at progressively deeper
levels of a synapse, leading to a clear temporal separation between long- and short-term memory. The
upper levels are more vulnerable to 'noise', i.e., the regular influx of (usually irrelevant) information which
is responsible for the phenomenon of forgetting; only short-term memories can be stored here. (For a
review of noise in neural systems, see [3]). Deeper synaptic levels are more protected from this noise, and
thus able to retain the memory of applied signals for a much longer time. Drawing on these structural
ideas first proposed by Fusi et al. [4], the model of [2] provides a theoretical framework for the dynamics
of signal propagation within the metaplastic synapse. It suggests that random signals are typically stored
in the upper levels of the synapse for relatively short times, and then lost to noise: non-random signals,
on the other hand, are stored as long-term memories in the deepest synaptic levels and forgotten much
more slowly.
The link between this theory and the experiment reported in [1] relies on the fact that the output
signal in the latter was amplified after a process of tetanic stimulation. This suggested the following
scenario: the initial action potentials, interpreted as a non-random signal, cumulatively built up a long-
term memory of the signal in the deepest synaptic levels. The synapse dynamics were then frozen
so that further discharge was prevented. When a further action potential was applied, the synaptic
dynamics restarted ('use'-dependence): the release of the accumulated memory from the deepest levels
of the synapse constituted the observed enhancement of the output signal described in [1]. While this
enhancement is plausibly accounted for by the model of metaplastic synapses [2], the explanation of
the freezing of the synaptic dynamics and its subsequent use-dependence needs the introduction of a
biological switch. The stochastic and bistable switch presented in this paper meets this need, and models
the role of kinase (CaMKII) in the actual experiment [1].
In the following, we first review the basics of the model of a metaplastic synapse [2]. We then provide
a full theoretical framework for the explanation of the experimental results, with an emphasis on the
dynamics of the biological switch. We close by discussing our results.
Results
A model of a metaplastic synapse
If synapses are highly plastic, signals are quickly stored: however, high plasticity also means that more
and more signals are stored, generating enough 'noise' so that 'memories' of earlier signals soon become
irretrievable. Clearly, this is at variance with the fact that long-term memories are ubiquitous in human
experience; it was to resolve this paradox that models of metaplastic synapses were formulated [4]. The
idea behind such models was that the introduction of 'hidden states' for a synapse would enable the
delinking of memory lifetimes from instantaneous signal response: while maintaining quick learning, this
mechanism would also be able to allow slow forgetting. This was implemented by the storage of memories
at different 'levels': the relaxation times for the memories increased as a function of depth. This hierarchy
of time scales models the phenomenon of metaplasticity [5, 6].
In [2] these ideas were put into a new framework, with the dynamics of signal processing playing a
central role. Also, two different internal synaptic structures were investigated: the first (Model I) was
3
very similar to Fusi's original model [4], while the second (Model II) had a different architecture. In the
following, we focus on the second model. We start by reviewing its essential features.
The dynamics of the model are defined in Figure 2. At every discrete time step t, the synapse is
subjected either to a potentiating pulse (PP) (encoded as ε(t) = +1) or to a depressing pulse (DP)
(encoded as ε(t) = −1), where ε(t) = ±1 is the instantaneous value of the input signal at time t. There
are three outcomes of the application of a PP signal:
• If the synapse is in its − state at depth n, it may climb one level (n → n − 1) with probability αn.
• If it is in its − state at depth n, it may alternatively cross over to the + state at the same level with
probability βn.
• If it is already in its + state at depth n, it may fall one level (n → n + 1) with probability γn.
The level-resolved output signal of level n at time t:
and the total output signal at time t:
Dn(t) = Qn(t) − Pn(t)
D(t) = Xn≥0
Dn(t)
(1)
(2)
can be expressed in terms of the probabilities Pn(t) (resp. Qn(t)) for the synapse to be in the − state
(resp. in the + state) at level n = 0, 1, . . . at time t = 0, 1, . . . The dynamical equations obeyed by the
latter probabilities are reviewed in the Methods section, along with other details for the mathematically
inclined reader.
Before any meaningful signal is applied, the synapse is assumed to be in its default state. The latter
state is defined as the stationary state reached by the synapse if subjected to a long random input signal.
It is described in detail in the Methods section (see equations (16) to (21)). When a single potentiating
pulse signal is applied at time t = 1 (that is, ε(1) = +1) to the synapse in its default state, the synapse
will get polarized in response, and thus 'learn' the signal. Later on, under the influence of a random input
signal for times t ≥ 2, it will 'forget' the PP signal, and return to its default state. Figure 3 shows plots
of the reduced output signal D(t)/D(1) against time t for several values of the control parameter β. All
subsequent figures refer to the parameter values β = 0.2, γ = 0.5, and ξs = ξd = 5 (see Methods section).
From here on, we will refer to times where the synapse is subjected to a significant signal (ε(t) = ±1)
as learning phases, and to times where the synapse is subjected to random input (ε(t) = 0) as forgetting
phases.
The late stages of the forgetting process are characterized by a universal power-law decay of the
output signal:
D(t) ∼ t−θ.
This is known as power-law forgetting [7 -- 9]. The forgetting exponent
θ = 1 +
ξd
ξs
(3)
(4)
is always larger than unity and depends on the ratio of the dynamical and static lengths ξd and ξs. If
the synapse were finite rather than infinite, and consisted of N levels, the power-law decay (3) would be
exponentially cut off at a time
τN ∼ exp(N/ξd)
(5)
which grows exponentially fast with the ratio of the number N of levels to the dynamical length ξd.
We now describe the effect of a sustained input of potentiating pulses lasting for T consecutive time
steps (ε(t) = +1 for 1 ≤ t ≤ T ) on the model synapse:
in the following, this will be referred to as a
long-term potentiating (LTP) signal. The synapse is again assumed to be initially in its default state.
The learning and forgetting processes are qualitatively similar to the PP case described above, while novel
4
qualitative features emerge when the duration of the signal is long enough (βT ≫ 1). In this regime,
the synapse gets almost totally polarized under the persistent action of the input signal. This saturation
phenomenon is illustrated in Figure 4, which shows the output signal D(t) for several durations T of the
LTP signal.
The synapse slowly builds up a long-term memory in the presence of a long enough LTP signal, as
the memorized signal moves to deeper and deeper levels. At the end of the learning phase (t = T ), the
polarisation profile will have the form of a sharply peaked traveling wave, around a typical depth which
grows according to the logarithmic law
n(T ) ≈ ξd ln γT.
(6)
The total output signal then decays according to the universal power law (3), irrespective of the duration
of the learning phase, driving home the universality of power-law forgetting.
The above model provides a mechanism for the long-term memory storage due to use dependence in
the Lymnaea synapse examined in [1]. With the tetanic stimulation acting as an LTP signal, this model
shows that the synapse becomes fully polarized, the memory of the stimulating pulses being 'stored'
in a deep level. However, another concept is needed to model the subsequent freezing of the synaptic
dynamics: a major clue is provided by the fact that it is suggested in [1] that the activation of CaMKII
in the presynaptic cell acts like a 'molecular switch'. Accordingly, we present a theoretical model of a
stochastic and bistable switch in the following section.
The effect of use-dependent synaptic potentiation:
coupling to a stochastic and bistable switch
In this section, we present a model of a biological switch to describe the role of CaMKII in the experiment
reported in [1]. The switch can exist in two states, 'on' or 'off', which we label by the binary variable
σ(t) = 1 or 0. Since natural processes are usually stochastic rather than deterministic, we incorporate this
by postulating that the switch is on with probability Π(t), and off with the complementary probability
1 − Π(t). Thus:
σ(t) = (cid:26)1 with probability Π(t),
0 with probability 1 − Π(t).
(7)
The main effect of this switch is to freeze the synaptic dynamics after adequate potentiation: we
accordingly refer to the probability Π(t) as the freezing probability. Thus:
• If the switch is off (σ(t) = 0), the synapse evolves as usual. This occurs with probability 1 − Π(t).
• If the switch is on (σ(t) = 1), the forgetting process ('discharge') is frozen. This occurs with probability
Π(t).
More precisely, the synapse learns via (12)-(13) and forgets via (18) when the switch is off. When the
switch is on, the synapse still learns via (12)-(13), but it does not forget at all.
In the experiment, a minimum of seven action potentials is needed to activate the switch and freeze
the dynamics: this suggests that the switch somehow responds to the saturation of the synaptic capacity,
so that freezing never sets in for less than this number of action potentials. Also, a further tetanic pulse
after a period of quietude is needed to restart the synaptic dynamics. We design the evolution of the
freezing probability Π(t) accordingly:
• If the synapse is within a learning phase (ε(t) = ±1 and ε(t − 1) = ±1), the freezing probability evolves
according to the quadratic rule
Π(t) = 1 − c(1 − Π(t − 1))2.
(8)
This rule ensures that the freezing probability increases with the number of action potentials applied,
saturating quickly to Π = 1 and freezing the synaptic dynamics after a threshold number of these is
reached. It is desirable that the fixed point of Π = 1 is superstable (see below), and the quadratic law (8)
is the simplest non-linear law which ensures this.
5
• If the synapse is in a forgetting phase (ε(t) = 0), the freezing probability is itself frozen to its value
inherited from the past:
Π(t) = Π(t − 1).
(9)
This rule ensures that once frozen, the dynamics stay frozen and that the synapse stays potentiated, until
a further action potential is applied.
• If the synapse is at the first step of a learning phase (ε(t) = ±1 but ε(t − 1) = 0), the freezing
probability is instantly reset to
Π(t) = 0.
(10)
This ensures that the synaptic dynamics restart, as soon as an action potential is applied for the first
time following a period of forgetting.
Note that there is a 'soft' threshold in the experiment for the switch to kick in, in that a minimum of
seven action potentials is required; as reported in [1], this phenomenon was usually seen to occur across
a range of 7 to 14 action potentials. The freezing probability which models the switch dynamics needs
to incorporate this soft threshold, for which superstability of the fixed point Π = 1 is desirable. This
ensures that at the end of a sustained LTP signal of duration T , Π(T ) converges very rapidly to unity
(more rapidly than exponentially), as soon as T exceeds a characteristic time T0, defined operationally
by Π(T0) = 1
2 . The value of the characteristic time fixes the parameter c. We have then
Π(T ) = 1 − exp(cid:18)−
2T −1 − 1
2T0−1 − 1
ln 2(cid:19) .
(11)
Figure 5 shows a plot of the freezing probability Π(T ) at the end of an LTP signal of duration T for
several values of the characteristic time T0 in the realistic range of 5 ≤ T0 ≤ 9. In practice, as soon as
Π(T ) is appreciably large, the switch kicks in stochastically -- and this can occur, as shown in Figure 5,
over a range of signal durations.
We now commence a global interpretation of the experiment. The synapse is assumed to be initially
in its default state, with Π(0) = 0. It is then subjected to a sustained LTP signal of duration T1 (i.e.,
the application of T1 action potentials), and to a single action potential at a much later time (T2 ≫ T1).
The synapse is subjected to a random input at all the other instants of time (ε(t) = +1 for 1 ≤ t ≤ T1
and for t = T2, else ε(t) = 0).
In the regime where the number of action potentials T1 of the initial signal is larger than the charac-
teristic time T0 of the switch, the freezing probability Π(T1) at the end of the LTP period is very high,
i.e., very close to unity (see Figure 5). During this learning phase, the output signal D(t) grows progres-
sively from D(0) = 0 to a large value D(T1). The high value of Π(T1) at the end of this phase typically
freezes the synaptic dynamics, ensuring that this enhanced output signal is not discharged. When the
next action potential is applied at time T2, the switch is turned off, and the synapse then relaxes via the
full discharge of the stored, enhanced output signal.
We now compare theory with experiment. Figure 6 shows a comparison between our theoretical
predictions (upper panel) with sharp-electrode electrophysiology recordings of a VD4/LPeD1 synaptic
pair (two lower panels). The theoretical prediction is meant to describe the average over many ensembles,
while the experimental data are assumed to be typical. Since the experimental averages are well-behaved
rather than subject to large fluctuations [1], a typical experimental output is representative of its average.
On the theoretical side, the time unit for the discrete updates is consistently chosen to be the time
interval between two successive applied APs, i.e., 100 ms. The black theoretical curve corresponds to 3
APs triggered during tetanic stimulation, which are insufficient to result in potentiation of a subsequent
excitatory postsynaptic potential (EPSP) in the LPeD1 neuron (T1 = 3 ≪ T0, and so the switch remains
off). The red theoretical curve corresponds to 11 APs, resulting in a potentiated response (T1 = 11 ≫ T0,
so the switch is turned on and the synapse is frozen). The predicted pattern of peak heights (symbols)
is rather robust, i.e., insensitive to model parameters, and provides a good overall description of the
experimental recordings.
6
The last part of this section reflects the 'bistability' of the switch, i.e., its stability in one of two
If the duration T1 of the initial LTP signal is comparable to the characteristic time T0, the
states.
synapse may exhibit two types of temporal behavior. It may behave as above for T1 = 11, i.e., stay
frozen until the subsequent action potential unfreezes it at time T2. This occurs with probability Π(T1)
(see Figure 5). The synapse may also remain unfrozen, so that the output signal decays almost entirely
before the subsequent action potential sets in. This occurs with the complementary probability 1 −Π(T1).
This bistable behavior is illustrated in Figure 7. We see from this figure that the difference between T1
and T0 provides a tunable parameter which determines how frequently the phenomenon described in
the experiment in [1] is observed: in the upper panel the parameters are such that it is almost always
observed, whereas in the lower panel it makes a much more random appearance. This parametrisation
is useful, since it can form the basis of future experiments. If we assume that the characteristic time T0
is, say, inversely proportional to the concentration of kinase, this suggests that a threshold concentration
can be identified for which a given number of action potentials will cause the short-term potentiation of
the synapse.
Discussion
Our theoretical model of a biological switch coupled to a metaplastic synapse is able to provide an
interesting framework for the interpretation of the experimental results of [1]. This novel form of synaptic
potentiation exhibits characteristics of short-term potentiation, but also exhibits characteristics of long-
term potentiation based on its time frame. While the enhanced output signal observed in the latter is
attributable to long-term potentiation of the synapse via tetanic stimulation, the use-dependence of the
synapse is explainable by the biological switch.
We start with the depotentiated state of the synapse, where, experimentally, single action potentials
triggered in VD4 elicit EPSP in LPeD1 of a non-potentiated amplitude. The application of repeated
action potentials in the pre-synaptic cell however builds up long-term memory in the synapse, i.e., a
potentiated EPSP in LPeD1. Meanwhile, during the tetanic stimulation, the freezing probability of the
molecular switch (CaMKII) evolves according to (8) in response to the growing saturation of the synapse.
If the number of action potentials is large enough, the switch is activated and synaptic dynamics are
fully frozen. Synaptic dynamics are only restarted when the switch is turned off: this happens when the
synapse is next 'used', and leads to the observed rapid decay back to a non-potentiated amplitude in
LPeD1. Thus, the critical dependence of the onset of the decay on synapse use, rather than the time
elapsed after priming, is explained by the model presented in this paper.
It is useful to discuss some of the key assumptions made here before concluding. The level of descrip-
tion of our model is sufficiently abstract that it does not claim to replicate intricate biological detail;
its description is limited to probing the unusually long times of use-dependent synaptic potentiation in
the Lymnaea synapse. From this point of view, the use of a model of long-term memory in metaplastic
synapses is perfectly appropriate. This model embodies the idea that forgetting takes place via the expo-
sure of the multi-level synapse to noise, from which the lowest synaptic levels are protected. This noise
could refer to fluctuations in neural/synaptic activity or to background signals; what is clear is that it is
ubiquitous and unavoidable. The model of a biological switch which we introduce depends both on the
presynaptic cell and on the state of the synapse: it is turned on when the synapse is saturated in response
to a series of action potentials, leading to the freezing of synaptic dynamics. It is turned off when the
synapse is next 'used', i.e an activation potential is applied after an inert period. An enhanced output
signal results, because of the discharge of accumulated memory at the synapse. The turning on and off of
the switch is stochastic, rather than deterministic, given the nature of most natural processes: and it is
bistable, since it can lead to two kinds of temporal behaviour at the synapse (with different probabilities,
of course). Finally, since the experimental recordings do not show strong fluctuations [1], the output of
a typical experiment can be meaningfully compared with the average output signal given by theory.
7
While the intact brain is composed of a significantly greater number of synapses with far more intricate
connectivity patterns, the current model has taken a reductionist approach to understanding synaptic
plasticity at the level of a single synapse. While previous forms of synaptic potentiation have been
modeled, this novel form of use-dependent synaptic potentiation has not. Therefore, this study may not
only be an important step in developing more complex models composed of multiple synapses, but also
be important in guiding further research in understanding neuronal network function in intact Lymnaea
brain.
Methods
The basic quantities used to describe the state of the synapse are the probabilities Pn(t) (resp. Qn(t))
for the synapse to be in the − state (resp.
in the + state) at level n = 0, 1, . . . at time t = 0, 1, . . .
These probabilities obey the following coupled linear equations, whose form is characteristic of Markov
chains [10]:
• ε(t + 1) = +1 (see Figure 2, left):
Pn(t + 1) = (1 − αn − βn)Pn(t) + αn+1Pn+1(t),
Qn(t + 1) = (1 − γn)Qn(t) + γn−1Qn−1(t) + βnPn(t).
• ε(t + 1) = −1 (see Figure 2, right):
Pn(t + 1) = (1 − γn)Pn(t) + γn−1Pn−1(t) + βnQn(t),
Qn(t + 1) = (1 − αn − βn)Qn(t) + αn+1Qn+1(t).
(12)
(13)
The transition probabilities of the model are assumed to decay exponentially with level depth n:
αn = αe−(n−1)µd ,
βn = βe−nµd ,
γn = γe−nµd ,
where the dynamical length
ξd =
1
µd
(14)
(15)
measures the number of fast levels at the top of the synapse.
The default state of the synapse is defined as its stationary state in the presence of a random input
signal, defined by choosing at each time step
ε(t) = (cid:26)+1 with probability 1
2 ,
−1 with probability 1
2 .
(16)
The dynamics of the synapse subjected to such a random input, referred to as a 'white-noise' random
input in [2], is defined by averaging the linear dynamical equations (12)-(13) over both instances of ε(t)
at each time step. The resulting average dynamics is formally labeled as ε(t) = 0.
It has a simpler
expression in terms of the sums and differences
namely
Sn(t) = Pn(t) + Qn(t), Dn(t) = Pn(t) − Qn(t),
Sn(t + 1) = Sn(t) + 1
Dn(t + 1) = Dn(t) + 1
2(cid:0)γn−1Sn−1(t) − (αn + γn)Sn(t) + αn+1Sn+1(t)(cid:1),
2(cid:0)γn−1Dn−1(t) − (αn + 2βn + γn)Dn(t) + αn+1Dn+1(t)(cid:1).
The default state of the synapse, defined as the stationary state of the above average dynamics, is
characterized by the probabilities
P st
n = Qst
n = 1
2 (1 − e−µs)e−nµs ,
(19)
(17)
(18)
with
µs = ln
α
γ
.
8
(20)
The default state is appropriately featureless and unpolarized, as it should be for a symmetric synapse.
The corresponding static length
ξs =
1
µs
(21)
gives a measure of the effective number of occupied levels in the default state.
Acknowledgments
A. M. thanks the Institut de Physique Th´eorique, where much of this work was carried out, for its
customary gracious hospitality during her visits.
References
1. Luk CC, Naruo H, Prince D, Hassan A, Doran SA, et al. (2011) A novel form of presynaptic
CaMKII-dependent short-term potentiation between Lymnaea neurons. Eur J Neurosci 34: 569-
577.
2. Mehta A, Luck JM (2011) Power-law forgetting in synapses with metaplasticity. J Stat Mech
P09025.
3. Faisal AA, Selen LPJ, Wolper DM (2008) Noise in the nervous system. Nature Reviews Neurosci
9: 292-303.
4. Fusi S, Drew PJ, Abbott LF (2005) Cascade models of synaptically stored memories. Neuron 45:
599-611.
5. Abraham WC, Bear MF (1996) Metaplasticity: The plasticity of synaptic plasticity. Trends in
Neurosciences 19: 126-130.
6. Fischer TM, Blazis DEJ, Priver NA, Carew TJ (1997) Metaplasticity at identified inhibitory
synapses in aplysia. Nature 389: 860-865.
7. Wixted JT, Ebbesen EB (1991) On the form of forgetting. Psychol Science 2: 409-415.
8. Wixted JT, Ebbesen EB (1997) Genuine power curves in forgetting: a quantitative analysis of
individual subject forgetting functions. Mem Cognit 25: 731-739.
9. Brown MW, Xiang JZ (1998) Recognition memory: neuronal substrates on the judgement of prior
occurrence. Prog Neurobiol 55: 149-189.
10. van Kampen NG (1992) Stochastic Processes in Physics and Chemistry. Amsterdam: North-
Holland.
Figure Legends
9
Figure 1. A representative electrophysiology trace showing the potentiation of the VD4/LPeD1
synapse following tetanic stimulation and its subsequent depotentiation. A tetanic stimulation of
presynaptic neuron VD4 (at thin arrow) results in a compound excitatory postsynaptic potential
(EPSP) in postsynaptic neuron LPeD1. A subsequently triggered action potential in VD4 (asterisk)
results in an EPSP of greater amplitude than pre-tetanic stimulation (thick arrow).
β
0
β
1
β
2
α
1
α
2
β
0
β
1
β
2
α
1
α
2
γ
0
γ
1
γ
0
γ
1
ε = +1 (LTP)
ε = −1 (LTD)
Figure 2. Schematic representation of the model of the internal synaptic structure. Arrows denote
possible transitions in the presence of a potentiating pulse (PP, ε = +1, left panel) and of a depressing
pulse (DP, ε = −1, right panel). Corresponding transition probabilities are indicated. In each panel, the
left (resp. right) column corresponds to the − (resp. +) state. The model studied in this work is
actually infinitely deep. After [2].
1.2
)
1
(
D
/
)
t
(
D
0.8
0.4
0.05
0.1
0.15
0.2
0
0
20
40
t
60
80
Figure 3. Plot of the reduced output signal D(t)/D(1) after a single PP input signal, against time t,
for several β. After [2].
10
1
0.8
0.6
0.4
0.2
)
t
(
D
1
3
10
30
0
0
20
40
60
80
100
t
Figure 4. Plot of the output signal D(t) against time t, for several durations T of the LTP signal.
After [2].
T0=5
T0=7
T0=9
1
0.8
0.6
0.4
0.2
)
T
(
Π
0
0
2
4
6
8
10
12
14
T
Figure 5. Plot of the freezing probability Π(T ) at the end of an LTP signal, against its duration T , for
three values of the characteristic time T0. The threshold for freezing is least for the black curve
(T0 = 5), and most for the blue (T0 = 9): while, say, 8 action potentials will definitely cause the onset of
freezing for the black curve (Π(8) = 0.997), it will only rarely do so for the blue one (Π(8) = 0.292).
11
)
t
(
D
1
0.8
0.6
0.4
0.2
0
0
3 APs
11 APs
20
40
60
t
80
100
120
140
Figure 6. An integrative figure showing the good qualitative agreement of theory and experiment. The
two lower panels show sharp-electrode electrophysiology recordings of a VD4/LPeD1 synaptic pair.
During the induction phase, three action potentials were triggered at ∼ 10 Hz. The synapse was allowed
to remain quiescent for ∼ 5 s and when a subsequent action potential was triggered, the amplitude of
the postsynaptic potential was similar to pre-tetanic stimulation. However, when eleven action
potentials were triggered at ∼ 10 Hz, a potentiated response was observed after the same quiescent
period of ∼ 5 s after stimulation. The upper panel shows the predictions of the theoretical model, also
for three and eleven action potentials.
12
0.7
0.6
0.5
0.4
0.3
0.2
0.1
)
t
(
D
T1=7 > T0=5
unfrozen
frozen
0.7
0.6
0.5
0.4
0.3
0.2
0.1
)
t
(
D
T1=T0=5
unfrozen
frozen
0
0
20
40
60
80
100
t
0
0
20
40
60
80
100
t
Figure 7. Plot of the two possible kinds of output signals D(t) generated by the protocol described in
the text, against time t (T0 = 5, T2 = 50). Symbol sizes are proportional to the probabilities of each
kind of behavior, i.e., Π(T1) for the frozen one and 1 − Π(T1) for the unfrozen one. Left: T1 = 7 is
larger than T0 = 5, so that Π(T1) = 0.946 is very high. Right: T1 = T0 = 5, so that Π(T1) = 1
2 .
|
1707.01373 | 2 | 1707 | 2017-12-15T12:59:47 | Looping and Clustering model for the organization of protein-DNA complexes on the bacterial genome | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.SC"
] | The bacterial genome is organized in a structure called the nucleoid by a variety of associated proteins. These proteins can form complexes on DNA that play a central role in various biological processes, including chromosome segregation. A prominent example is the large ParB-DNA complex, which forms an essential component of the segregation machinery in many bacteria. ChIP-Seq experiments show that ParB proteins localize around centromere-like parS sites on the DNA to which ParB binds specifically, and spreads from there over large sections of the chromosome. Recent theoretical and experimental studies suggest that DNA-bound ParB proteins can interact with each other to condense into a coherent 3D complex on the DNA. However, the structural organization of this protein-DNA complex remains unclear, and a predictive quantitative theory for the distribution of ParB proteins on DNA is lacking. Here, we propose the Looping and Clustering (LC) model, which employs a statistical physics approach to describe protein-DNA complexes. The LC model accounts for the extrusion of DNA loops from a cluster of interacting DNA-bound proteins. Conceptually, the structure of the protein-DNA complex is determined by a competition between attractive protein interactions and the configurational and loop entropy of this protein-DNA cluster. Indeed, we show that the protein interaction strength determines the "tightness" of the loopy protein-DNA complex. With this approach we consider the genomic organization of such a protein-DNA cluster around a single high-affinity binding site. Thus, our model provides a theoretical framework to quantitatively compute the binding profiles of ParB-like proteins around a cognate (parS) binding site. | physics.bio-ph | physics |
Looping and Clustering model for the organization of protein-DNA complexes on the
bacterial genome
Jean-Charles Walter, Nils-Ole Walliser, Gabriel David, J´erome Dorignac, Fr´ed´eric Geniet, and John Palmeri
Laboratoire Charles Coulomb, UMR5221 CNRS-UM, Universit´e de Montpellier,
Place Eug`ene Bataillon, 34095 Montpellier Cedex 5, France
Andrea Parmeggiani
DIMNP, UMR5235 CNRS-UM, Universit´e de Montpellier,
Place Eug`ene Bataillon, 34095 Montpellier Cedex 5, France
Department of Molecular Biology and Lewis-Sigler Institute for Integrative Genomics,
Princeton University, Princeton NJ 08544, USA
Ned S. Wingreen
Chase P. Broedersz∗
Arnold-Sommerfeld-Center for Theoretical Physics and Center for NanoScience,
Ludwig-Maximilians-Universitat Munchen, D-80333 Munchen, Germany.
(Dated: December 18, 2017)
The bacterial genome is organized by a variety of associated proteins inside a structure called
the nucleoid. These proteins can form complexes on DNA that play a central role in various bio-
logical processes, including chromosome segregation. A prominent example is the large ParB-DNA
complex, which forms an essential component of the segregation machinery in many bacteria. ChIP-
Seq experiments show that ParB proteins localize around centromere-like parS sites on the DNA to
which ParB binds specifically, and spreads from there over large sections of the chromosome. Recent
theoretical and experimental studies suggest that DNA-bound ParB proteins can interact with each
other to condense into a coherent 3D complex on the DNA. However, the structural organization of
this protein-DNA complex remains unclear, and a predictive quantitative theory for the distribution
of ParB proteins on DNA is lacking. Here, we propose the Looping and Clustering (LC) model,
which employs a statistical physics approach to describe protein-DNA complexes. The LC model
accounts for the extrusion of DNA loops from a cluster of interacting DNA-bound proteins that is
organized around a single high-affinity binding site. Conceptually, the structure of the protein-DNA
complex is determined by a competition between attractive protein interactions and the configura-
tional and loop entropy of this protein-DNA cluster. Indeed, we show that the protein interaction
strength determines the "tightness" of the loopy protein-DNA complex. Thus, our model provides
a theoretical framework to quantitatively compute the binding profiles of ParB-like proteins around
a cognate (parS) binding site.
I.
INTRODUCTION
Understanding the biophysical principles that govern
chromosome structure in both eukaryotic and prokary-
otic cells remains an outstanding challenge [1–7]. Many
bacteria have a single chromosome with a length three
orders of magnitude longer than the cell itself, posing a
daunting organizational problem. Owing to recent tech-
nological advances in live-cell imaging and chromosome
conformation capture based approaches, it is becoming
increasingly clear that the DNA is not coiled like a sim-
ple amorphous polymer inside the cell [8–10], but rather
exhibits a high degree of organization over a broad range
of lengthscales [11]. It remains unclear, however, how this
spatial and dynamic organization of the chromosome is
established and maintained inside living bacteria [12]. A
∗Electronic address: [email protected]
host of Nucleoid-Associated Proteins (NAPs) have been
shown to play a central role in the spatial organization
of the bacterial chromosome [12–14]. Such NAPs bind
to the DNA in large numbers, and by interacting with
each other and with DNA in both sequence-dependent
and sequence-independent manners they can collectively
structure the DNA polymer and control chromosome or-
ganization.
In many bacterial species,
the broadly conserved
ParABS system is responsible for chromosome and plas-
mid segregation [12, 15]. A central component of this
system is the partitioning module, which is formed by
a large protein-DNA complex of ParB proteins that as-
sembles around centromere-like parS sites, frequently lo-
cated near the origin of replication. The ParBS com-
plexes can subsequently interact with ParA ATPases,
leading to the segregation of replicated origins [16–23].
How is this ParBS partitioning module physically orga-
nized on the DNA? ParB is known to bind specifically
to parS, triggering the formation of a large protein-DNA
cluster, which is visible as a tight focus in microscopy
2
model with short range interactions. On general statis-
tical physical grounds, such a 1D model cannot be ex-
pected to account for the formation of a large coherent
protein-DNA complex, given physiological protein inter-
action strengths [29]. Furthermore, the number of ParB
proteins available in the cell is not sufficient to allow
enrichment by simple 1D polymerization of ParB along
DNA at genomic distances from parS as large as observed
experimentally [25]. To resolve the puzzle of how ParB
proteins organize around a parS site, we recently intro-
duced a novel theoretical framework to study the collec-
tive behavior of interacting proteins that can bind to a
DNA polymer [29]. This model suggested that ParB as-
sembles into a three-dimensional complex on the DNA,
as illustrated in Figure 1a,b. Single molecule experiments
provided direct evidence for the presence of 3D bridging
interactions between two ParB proteins on DNA [30, 31].
We showed that a combination of such a 3D bridging
bond and 1D spreading bonds between ParB proteins
constitutes a minimal model for the condensation of ParB
proteins on DNA into a coherent complex [29], consistent
with the observation that ParB-GFP fusion proteins form
a tight fluorescent focus on the DNA [15, 19, 24, 25].
The statistical properties of the 3D structure of ParB-
DNA complexes determines the binding profile of ParB
on DNA, which can be accurately measured in ChIP-Seq
experiments. However, it is computationally demanding
to simulate these binding profiles with the Spreading &
Bridging model. The protein binding profiles can be eas-
ily calculated analytically in the limit of strong protein-
protein interactions, where the cluster of ParB on the
DNA becomes compact with a corresponding triangu-
lar distribution of ParB along DNA. The protein binding
profiles can also be estimated in the limit of weak protein-
protein interactions with the so-called Stochastic Binding
model, where a sphere of high ParB concentration is as-
sumed to exist within which a DNA polymer freely fluc-
tuates [25] (see Figure 1c). The description of the average
protein binding profile is thus similar to the return statis-
tics of the polymer into the ParB sphere [32], suggesting
a long range (power-law) distribution of ParB proteins
along DNA. Importantly, however, neither of these two
existing approaches provide a simple way of computing
ParB binding profiles around parS sites over the full rele-
vant range of system parameters. In addition, it remains
unclear how the Spreading & Bridging model and the
Stochastic Binding model relate to each other.
Here, we propose a theoretical approach to describe
the distribution of ParB proteins around parS sites on
the DNA in terms of molecular interaction parameters
and protein expression levels. To this end, we develop
a simple model for protein-DNA clusters that explicitly
accounts for the competition between the positional en-
tropy associated with placing the loops on the cluster,
which favours a looser cluster configuration, and both
protein-protein interactions and loop closure entropy,
which tend to favour a compact cluster. This Looping
and Clustering model represents a reduced, approximate
FIG. 1:
Schematic illustration of two recent models pro-
posed to describe the ParB partition complex (left) accom-
panied with a typical distribution of ParB on extended DNA
(middle), and the average distribution profile (right). The
Spreading & Bridging model [29] is shown with (a) strong
coupling JS → ∞, where thermal fluctuations cannot break
the bonds between proteins such that all bridging and spread-
ing interactions are satisfied, and (b) intermediate coupling
where the energetic cost of breaking a spreading bond com-
petes with the configurational and loop entropy. With the
Looping and Clustering approach presented here, we propose
a simple analytic description for this regime. (c) The Stochas-
tic Binding model assumes a spherical region of high concen-
tration of ParB around parS [25]. This model can be seen
as taking the limit of the spreading bond strength to zero
(JS → 0), and thus the formation of loops is not hampered
by protein-protein bonds.
In this limit, the binding profile
can be described as the return of the polymer to an origin of
finite size, such that the profile is given by P (s) ∝ (s + C)−dν ,
where d is the dimension, ν is the Flory exponent, and C is a
constant.
images of fluorescently labeled ParB [15, 19, 24, 25]. The
propensity of ParB to form foci around parS has been
exploited in recent studies, which used exogenous ex-
pression of fluorescently labeled ParB along with parS
insertion to label DNA loci for live-cell imaging [26, 27].
In the F-plasmid of Escherichia coli cells, each ParB fo-
cus contains roughly 300 proteins, together representing
90% of all ParB present in the cell [25]. High-precision
ChIP-Seq experiments on this system provide quantita-
tive ParB binding profiles along the DNA, which are
strongly peaked around parS with a broad decay over
a distance of up to 13 kilobasepairs (kb), consistent with
earlier observations [24, 28].
Various models have been introduced to explain the
distribution of ParB along DNA around parS sites. An
early study of the distribution of ParB proposed that
ParB proteins spread from the parS sequence by nearest-
neighbor interactions, forming a continuous filament-like
structure along the DNA [28]. This model was termed
the Spreading model. However, this is effectively a 1D
(a)(b)(c): parS: DNA outside the cluster: DNA inside the cluster: ParB proteinP(s)sP(s)sP(s)~[s+C]-1.5J =J = 0SS0011parSparSP(s)s?Intermediate JS0parS: ParB bound to parSJS=1JS=0JS(a)(b)(c): parS: DNA outsidethecluster: DNA insidethecluster: ParB proteinP(s)sP(s)sP(s)~[s+C]-1.5J =J =0SS0011parSparSP(s)s?IntermediateJS0parS: ParB boundtoparS(a)(b)(c): parS: DNA outsidethecluster: DNA insidethecluster: ParB proteinP(s)sP(s)sP(s)~[s+C]-1.5J =J =0SS0011parSparSP(s)s?IntermediateJS0parS: ParB boundtoparSp(s)⇠(s+C) d⌫version of the full Spreading & Bridging model that incor-
porates the key physical ingredients needed to provide a
clearer understanding and at the same time greatly facil-
itates calculations of the distribution profile of ParB (or
other proteins that form protein-DNA clusters). Thus,
our approach can be used to estimate molecular interac-
tions between proteins from experimentally determined
protein binding profiles.
II. THE LOOPING AND CLUSTERING MODEL
To theoretically describe the protein binding profiles of
ParB on DNA, we first consider a DNA polymer of length
L that can move in space on a 3D cubic lattice and with
a finite number of proteins m. Since the number of ParB
proteins in the protein-DNA cluster has been observed
to include the vast majority of proteins in the cell [25],
we employ a canonical ensemble with a fixed number of
ParB proteins m in the ParB complex. These proteins
are able to diffuse along the DNA. Importantly, in this
model the DNA itself is also dynamic and fluctuates be-
tween different three-dimensional configurations, which
are affected by the presence of interacting DNA bound
proteins. When proteins are bound to the DNA, they
are assumed to be able to interact attractively with each
other by contact interactions in two distinct ways: (i) 1D
spreading interactions with coupling strength JS, defined
as an interaction between proteins on nearest-neighbor
sites along the polymer, and (ii) a 3D bridging interaction
with strength JB between two proteins bound to sites on
non-nearest neighbor-sites on the DNA, but which are
positioned at nearest neighbor-sites in 3D space (see Fig-
ure 1a,b). Thus, these bridging interactions couple to the
3D configuration of the DNA, while the 1D spreading in-
teractions do not. Single-molecule experiments provide
evidence for bridging bonds [30], with the bridging va-
lency of a ParB protein limited to one [33, 34]. Even
in this case where each protein can form two spreading
bonds and a single bridging bond, the system has been
shown to exhibit a condensation transition where the ma-
jority of the proteins form a single large cluster that can
be localized by a single parS site on the DNA [29].
While it is possible to perform Monte Carlo simulations
of the Spreading & Bridging model for a lattice poly-
mer, such simulations are computationally demanding.
In this paper, we aim to provide a simple analytical de-
scription for the average binding profile of proteins along
the DNA (see right panels in Figure 1). With this aim in
mind, we can simplify our description by realizing that
the configurations of ParB proteins along the DNA are
more sensitive to JS than to JB, for sufficiently large JB.
While both spreading and bridging bonds are necessary
for the condensation of all proteins into a single cluster,
loop extrusion from the cluster is controlled by JS, and
such loop extrusion strongly impacts the binding profile
of proteins on the DNA. Indeed, a loop can be extruded
from the protein-DNA cluster by breaking a spreading
3
bond, but without effecting the internal configuration of
the bridging bonds. Therefore, we will assume that JB is
sufficiently large to maintain a coherent 3D protein-DNA
cluster, leaving JS as the main adjustable parameter in
the model.
A contiguous 3D cluster of proteins on DNA with loops
can effectively be represented graphically by a discon-
nected 1D cluster along the DNA, where connections in
3D between the 1D subclusters are implied, and domains
of protein-free DNA within the disconnected 1D cluster
represent loops that emanate from the 3D cluster (see
Figure 1b,c). We can describe this system by a reduced
model for the effective 1D cluster in which we account for
the entropy of the loops that originate from the protein-
DNA cluster. In this model, the spreading bond energy
set by the parameter JS combined with the cost in loop
closure entropy, competes with the positional entropy for
placing loops on the cluster and will therefore play a cru-
cial role in determining the binding profile of ParB on
DNA around a parS site.
To capture these effects, we propose the reduced Loop-
ing and Clustering (LC) model, which offers a simplified
description of 3D protein-DNA clusters with spreading
and bridging bonds. In this model a loop is formed when-
ever there is a gap between 1D clusters. We can make the
connection between the gaps in the 1D cluster and the
number of loops extending from the 3D cluster explicit
by writing down the partition function for this model.
The effective 1D cluster corresponding to a 3D cluster
with m proteins and n loops has a multiplicity:
Ωcluster =
(m − 1)!
(m − n − 1)!n!
,
(1)
which counts the number of ways in which one can parti-
tion m proteins into n + 1 subclusters in 1D. This multi-
plicity leads to a positional entropy of mixing, Scluster =
ln Ωcluster, for placing n loops at m − 1 possible posi-
tions (in units of kB). Note, we do not explicitly include
the number of ways in which the bridging bonds can be
formed, since loop formation is not expected to substan-
tially affect the possible configurations of bridging bonds.
However, creating n loops will require breaking n spread-
ing bonds, and the probability at equilibrium for this
to occur will include a Boltzmann factor ∼ exp (−nJS),
where the interaction energy is expressed in units of kBT .
Within our simple description, we do not consider how
the formation of a loop affects the full internal entropy
of the protein-DNA cluster, but this can be expected to
be a fixed number per loop that can be absorbed into JS.
Furthermore, the loops that are formed are assumed to
be independent, and thus contribute to the loop closure
entropy (in units of kB) as [32]:
n(cid:88)
i=1
Sloop = −dν
ln((cid:96)i) ,
(2)
where d is the spatial dimension, ν is the Flory exponent,
and the loop length is measured in units of the lattice
spacing of the polymer a, which we take to be equal to the
footprint of a ParB protein, e.g. 16 bp for the exogenous
ParABS system of E. coli [25].
This entropy is obtained by considering both the loops
formed within the protein cluster and the protein-free
segment of DNA outside the cluster. Indeed, the number
of configurations associated with loop i for a Gaussian
polymer is given by z(cid:96)i(cid:96)−dν
[32, 35], where z is the lat-
tice coordination number. Therefore, there is also an ex-
tensive contribution to the entropy given by kB(cid:96)i log(z).
i
However, when a loop of length (cid:96)i forms, the same length
of polymer is removed from the DNA outside of the clus-
ter, which also results in a reduction of the entropy by
kB(cid:96)i log(z). Thus, there is a precise cancelation between
the extensive contribution to the entropy associated with
the loop inside the cluster and the extensive contribution
due to effectively shortening the DNA outside the clus-
ter [39].
It is now straightforward to write down the partition
function of the Looping and Clustering model:
4
m−1(cid:88)
(m − 1)!
n=0
(m − n − 1)!n!
[1 + exp(−J(cid:48)S)]m−1
(cid:90) (cid:96)max
(cid:96)0
(cid:90) (cid:96)max
(cid:96)0
d(cid:96)n(cid:96)−dν
n
d(cid:96)1(cid:96)−dν
1
...
exp(−nJS)
.
(3)
ZLC
=
−→(cid:96)max→∞
where J(cid:48)S = JS + ln(cid:2)(cid:96)dν−1
0
(dν − 1)(cid:3) > JS is a renor-
where Lj =(cid:80)j−1
malized loop activation energy that includes the cost in
loop closure entropy). All lengths are measured in units
of the protein's footprint a, (cid:96)0 is the lower cutoff of loop
sizes and approximately represents the persistence length
of DNA, and the bond interactions are in units of kBT .
In the partition function, we conveniently set the upper
boundary of integration, (cid:96)max, to infinity. Strictly speak-
ing, the upper boundary for (cid:96)j should be L − (m + Lj),
i=1 represents the total accumulated loop
length before loop j. In practice, however, for chromo-
somes, but arguably also for plasmids, L (cid:29) m and the
probability to have a large loop is very small. For in-
stance, if we consider the F-plasmid of E. coli with a
length of 60 kbp, we have L = 3750 in units of the ParB
footprint of 16 bp [25, 36]. For this system, Monte Carlo
(MC) simulations (see Appendix A) of the LC model,
with m = 100 reveal that the average cumulated loop
size is ≈ 500 for small couplings (JS = 1) down to ≈ 25
for large couplings (JS = 4), which in both cases is much
less than the DNA length. Thus, for biologically relevant
cases it is reasonable to assume that the length of the
DNA polymer is much larger than the footprint of the
whole protein complex on the DNA.
The LC model constitutes a simple statistical mechan-
ics approach to describe how proteins assemble into a
protein-DNA cluster with multiple loops. Next, we will
include a parS site on the DNA, to which ParB proteins
bind with a higher affinity than the other non-specific
binding sites on the DNA. Our central aim is to compute
the binding profile of ParB around this parS site.
III. PROFILE OF PAR B FOR FIXED NUMBER
AND SIZES OF LOOPS
With our approach, we aim to quantitatively describe
average ParB binding profiles, which are directly mea-
surable by ChIP-Seq experiments. By fitting our model
to such ChIP-Seq data, it would be possible to extract
microscopic parameters such as the number of proteins
in the ParB clusters and the protein-protein interaction
parameters such as JS. In this section, we will describe
how to compute the ParB binding profile around this
parS site given a fixed number of loops with specified
loop lengths. Then we will use the statistical mechanics
framework provided above, to perform a weighted aver-
age over all possible loop numbers and sizes to arrive at
a simple predictive theory for the ParB binding profile.
A. 1-loop binding profile
It is instructive to start our analysis of ParB binding
profiles by first calculating the probability of ParB occu-
pancy as a function of distance from the parS site for the
case of a protein-DNA cluster with only one DNA loop
(n = 1) with fixed loop length (cid:96). We will assume a fixed
number m of ParB proteins in this 1-loop protein-DNA
cluster, and that one of these proteins is bound to the
parS site at any time, as illustrated in Figure 2. Thus, to
calculate the 1-loop ParB binding probability, P1(s, (cid:96)),
at a distance s from parS, we need to consider all possi-
ble configurations of proteins in the protein-DNA cluster
subject to these constraints.
First, we note that P1(s, (cid:96)) = 0 for s > m + (cid:96), be-
cause the 1D cluster can maximally extend to a distance
m + (cid:96), which occurs when the 1D cluster adopts a con-
figuration that lies entirely on one side of the parS site.
For a binding site at a distance s < m + (cid:96), the ParB
5
extruding from the cluster, i.e. an unoccupied site on
the DNA within the protein cluster, as depicted in Fig-
ure 2. The overbar here represents the complementary
condition, and the expression above simplifies because
P1(s, (cid:96)loop(s)) = 0 by construction.
We can proceed to calculate the conditional probabil-
ity, P1(s, (cid:96)loop(s)), by decomposing this contributions
as a sum of probabilities of mutually exclusive configu-
rations, which are conditioned by the location s(cid:48) of the
right edge of the 1D ParB cluster denoted as "end(s(cid:48))"
(see Figure 2). Then, we will take a continuous limit
for the binding profile assuming m (cid:29) 1, and express the
binding profile P1(s, (cid:96)loop(s)) in terms of probabilities,
pend(s(cid:48)), for the condition describing the position of the
right edge of the cluster. Thus, we first write the con-
ditional probability P1(s, (cid:96)loop(s)) for s ≥ 0 (the case
s < 0 is obtained by symmetry) as:
m+(cid:96)(cid:88)
(cid:90) m+(cid:96)
P1(s, (cid:96)loop(s), end(s(cid:48)))pend(s(cid:48)),
s(cid:48)=0
(5)
P1(s, (cid:96)loop(s)) =
≈
ds(cid:48)Θ(s(cid:48) − s)pend(s(cid:48)).
0
Clearly, P1(s, (cid:96)loop(s), end(s(cid:48))) = 1 when s < s(cid:48) and
zero otherwise, and thus we have replaced this term by
the Heaviside step function Θ(s(cid:48) − s) and approximated
the sum by an integral in the second line above.
To calculate pend(s(cid:48)), it is convenient to introduce two
subclusters, 1 and 2, with m1 and m − m1 proteins re-
spectively (0 < m1 < m), such that cluster 1 with m1
proteins is overlapping with parS, as shown in Figure 2.
Given two such subclusters, two equally likely situations
can occur: (i) the leftmost cluster overlaps with parS, i.e.
m − m1 + (cid:96) ≤ s(cid:48) < m + (cid:96) or (ii) the rightmost cluster
overlaps with parS, i.e. 0 ≤ s(cid:48) < m1. This directly al-
lows us to construct the conditional probability to find
the right edge of the whole system, such that one of the
m1 proteins in the cluster overlaps with parS:
pend(s(cid:48)m1) = 1
2m1
[Θ(s(cid:48) − (m − m1 + (cid:96)))Θ(m + (cid:96) − s(cid:48))
(6)
+ Θ(m1 − s(cid:48))] ,
where the prefactor 1/2 comes from the equal probabili-
ties to find the system in one of the two cases (i) and (ii).
The conditions (i) and (ii) are encoded with a product of
two unit step functions for (i) and a single step function
for (ii). Each single realization can be obtained by shift-
ing the position of the site in cluster 1 overlapping with
parS and is equally likely, giving rise to an overall prefac-
tor 1/m1. From this, we can obtain the full probability
pend(s(cid:48)) by integrating over m1:
FIG. 2: Schematic of the system with m proteins and a single
loop of size (cid:96). The whole cluster is split in two parts: m1 is
the number of proteins in the cluster that overlaps with parS
and m − m1 is the number of proteins in the other cluster.
The origin of the genomic coordinates is parS, the right edge
of the system (RE) is located at the coordinate s(cid:48). We can
divide the configurations into two equally likely cases: (i) the
leftmost cluster overlaps with parS or (ii) the rightmost cluster
overlaps with parS.
FIG. 3: Protein occupation probability, P1(s, (cid:96)), for a site a
genomic distance s from the parS site for different loop lengths
(cid:96) and a fixed cluster size of m = 200 proteins. Solid curves
represent analytic calculations from Eqs. (4), (8), and (??),
and dashed curves represent data obtained from exact numer-
ical enumeration for comparison to our analytical approxima-
tions. We note that for (cid:96) = 0, we recover the triangular profile
of the S&B model in the strong coupling limit JS → ∞ [29].
binding probability is reduced, either by configurations
where this site is located on the DNA loop within the
1D cluster, or by states where the 1D cluster adopts a
configuration around the parS site that does not extend
to the binding site at s, placing this site outside the 1D
cluster. To capture these effects, it is helpful to express
P1(s, (cid:96)) in terms of conditional probabilities:
P1(s, (cid:96)) = P1(s, (cid:96)loop(s))ploop(s) + P1(s, (cid:96)loop(s))ploop(s)
(4)
= P1(s, (cid:96)loop(s))ploop(s),
where "loop(s)" represents a condition with probabil-
ity ploop(s) corresponding to site s being part of a loop
sparS0s'mm-msparS0s'mm-mREREcase (i)case (ii)ll1111``050100150200250300s00.20.40.60.81P1(s,l)l = 0l = 25l = 50l = 75l = 100050100150200250300s00.20.40.60.81P1(s,l)l=0l=25l=50l=75l=100`050100150200250300s00.20.40.60.81P1(s,l)l=0l=25l=50l=75l=100`````050100150200250300s00.20.40.60.81P1(s,l)l=0l=25l=50l=75l=100050100150200250300s00.20.40.60.81P1(s,l)l=0l=25l=50l=75l=100(cid:90) m−1
1
pend(s(cid:48)) ≈
dm1 pend(s(cid:48)m1)p(m1)
Θ(m + (cid:96) − s(cid:48))
m(m − 2)
=
+ Θ(1 − s(cid:48))(m − 2) + Θ(s(cid:48) − 1)Θ(m − 1 − s(cid:48))(m − 1 − s)]
[Θ(1 − (m + (cid:96) − s(cid:48)))(m − 2) + Θ(m + (cid:96) − s(cid:48) − 1)Θ(s(cid:48) − (cid:96) − 1)(s(cid:48) − (cid:96) − 1)
6
(7)
where we used p(m1) = 2m1/(m(m − 2)), since the
number of configurations to place cluster 1 is ∝ m1 and
m1 ∈ [1, m− 1]. Using this expression for the normalized
probability distribution for the right edge of the 1D clus-
ter to be positioned at s(cid:48), we can compute the conditional
probability in Eq. (5):
P1(s, (cid:96)loop(s)) ≈
Θ((cid:96) + m − s)
(m − 2)m
Θ((cid:96) − s + 1) + (m − 2)Θ((cid:96) + m − s − 1)
(cid:20) (m − 2)2
(cid:18) (m − 2)2
2
+ (m − 2)(m + (cid:96) − s)Θ(s − m − (cid:96) + 1) + (m + (cid:96) − s − 1)m − (cid:96) + s − 3)Θ(s − (cid:96) − 1)
+ Θ(1 − s)
Θ(m − s − 1) + (m − 2)(1 − s)
(s − m + 1)2
Θ(s − 1)Θ(m − s − 1)
+
2
2
(cid:19)
(cid:21)
(8)
To obtain the full 1-loop protein distribution (Eq. (4)),
we first need to compute the probability for a site to not
be part of loop,
ploop(s) = 1 − ploop(s).
(9)
loop(s) = (cid:96)ρuni(m, (cid:96)) = (cid:96)
If the loop density, ρ, were uniform, we would simply
have puni
m+(cid:96) , since the 1D cluster
has a total length of m + (cid:96) with a single loop of length (cid:96).
This uniform condition would only apply if we randomly
choose (cid:96) sites to be part of the loop and ignore the re-
quirement that all these loop sites need to be neighbor-
ing. In a real cluster, however, we expect the loop density
ρloop(s) to be higher in the bulk of the 1D cluster than
close to the parS site or the edges, because fewer loops
can be formed near the parS site or near the boundaries
of the 1D cluster, at which a protein must be bound by
construction. In particular, we expect the loop density,
ρ(s, m, (cid:96)) ∝ min(s, (cid:96), m+(cid:96)−s), which measures the num-
ber of ways a site at s can be part of a loop. This results
in the normalized probability:
ploop(s) = (cid:96)ρ(s, m, (cid:96))
Θ(m − (cid:96)) [(cid:96)2 + (cid:96)(m − (cid:96))] + Θ((cid:96) − m)(cid:0) m+(cid:96)
(cid:96) min(s, (cid:96), m + (cid:96) − s)
2
(cid:1)2
(10)
=
In the normalization of this expression we distinguish
the cases where the loop is either smaller or larger than
the number of proteins in the cluster. With Eqs. (8)
and (10), we have all the elements to calculate the 1-loop
protein binding profile P1(s, (cid:96)) from Eq. (4).
We investigated the binding profiles P1(s, (cid:96)) predicted
by this model for a selected set of parameters, as shown
in Figure 3. We only show s > 0 because of the sym-
metry of the binding profile. It is instructive to contrast
these profiles with the triangular profile (black curve) for
a cluster with no loops. As expected, the addition of
loops widens the profile, allowing the tail of the distribu-
tion to extend out to a distance m + (cid:96). The widening of
the binding profile is accompanied by a faster decay of
the profile in the vicinity of parS, which crosses over to a
flatter profile at distances s > (cid:96) due to additional contri-
butions from configurations where the loop lies between
the parS site and site s.
Interestingly, for large loop size the profile can even
become non-monotonic with a slight increase near the
far edges of the domain. These features of the profile
reflect the reduced loop density near parS and near the
far edges of the cluster. Note that the integral under
this curve remains constant for varying (cid:96) to conserve the
number of particles in the cluster. To verify the validity
of the analytical approximations leading to P1(s, (cid:96)), we
used exact enumeration as a benchmark. Overall, the nu-
merics (dashed lines) and the analytics (solid lines) are in
good agreement for the 1-loop case, as shown in Figure 3.
In the next section, we employ the approximate analyt-
ical expressions obtained above, to efficiently calculate
the full binding profile averaged over all configurations.
7
IV. PROTEIN BINDING PROFILES AND
STATISTICS OF THE LOOPING AND
CLUSTERING MODEL
Above we defined the Looping and Clustering model
and calculated the binding profile of proteins around a
parS site for a cluster with 1 loop with fixed length. Real
protein-DNA clusters, however, are expected to fluctuate
with new loops forming and disappearing continuously.
To capture such fluctuations, we will use the expressions
for the binding profile of a static cluster with fixed loop
length together with a statistical mechanics description
of the LC model to obtain average binding profiles for dy-
namic clusters, including an ensemble average over both
the number of loops and the loop lengths.
To obtain a full binding profile averaged over all real-
izations, it is useful to investigate the statistics of loops
that extend from the protein-DNA cluster and how these
statistics are determined by the underlying microscopic
parameters of the model. We start by considering the
number of loops that extend from the cluster. Using the
partition function in Eq. (3), it is possible to calculate
the basic features of the LC model. For instance, the
moments of the distribution of the number of loops are
given by:
(cid:104)nα(cid:105) =
1
ZLC
nα
(m − 1)!
(m − n − 1)!n!
e−nJS
(11)
From this, we find the the average loop number is given
by:
m−1(cid:88)
n=0
(cid:35)n
.
(cid:34)
(cid:96)1−dν
0
dν − 1
(cid:104)n(cid:105) = (m − 1)
1
1 + eJ(cid:48)
S
me−JS ,
(12)
∼m,J(cid:48)
S(cid:29)1
(a) Average number of loops, (cid:104)n(cid:105), as a function of
FIG. 4:
spreading coupling strength JS obtained from Eq. (12). The
different curves correspond to protein number m = 50 (black)
m = 100 (red), m = 200 (green), and m = 400 (blue), with
loop-size cutoff (cid:96)0 = 10. We observe an exponential decrease
(cid:104)n(cid:105) ∝ e−JS in accordance with Eq.(12).
Inset: Same data
replotted with expected dependence of average loop number
on m scaled out. (b) Average number of loops (cid:104)n(cid:105) as a func-
tion of m for JS = 1, 2, 3, and 4. The behaviour is linear
as expected from Eq.(12). The prefactor that determines the
vertical shift between the different curves scales with e−JS , as
demonstrated in the inset of panel (b). (c) Average loop prob-
ability as a function of the genomic coordinate with m = 200
and L = 4000 for protein-DNA clusters with fluctuating loop
number and loop lengths. Different curves correspond to dif-
ferent spreading couplings JS = 1, 2, 3, and 4. The analytic
approximation using Eq.(15) for the loop density, averaged
over different loop configurations with the appropriate Boltz-
mann factor as in Eq. (14) is compared to MC simulations
(dashed curves) of the LC model (see Appendix A).
where J(cid:48)S is the renormalized loop activation energy in-
troduced in Eq. (3). The average loop number (cid:104)n(cid:105) is
depicted in Figure 4a, demonstrating the exponential de-
pendence on the spreading energy JS. In Figure 4b, we
plot (cid:104)n(cid:105) as a function of the total number of proteins m
in the protein-DNA cluster. Indeed, we observe the ex-
pected linear dependence of the average loop number (cid:104)n(cid:105)
on m over a broad range of parameters. These results
illustrate how the average number of loops is determined
by the competition between the effective renormalized
loop activation energy, J(cid:48)S (including the cost in loop
closure entropy), and the gain in the positional entropy
of mixing (see Appendix B).
The linear dependence on m in Eq. (12) reflects that
loops are assumed to be able to form anywhere in the
cluster in the Looping and Clustering model. However,
one would naively expect that loops can only form at
the surface of a 3D cluster, resulting in a dependence
(cid:104)n(cid:105) ∼ m2/3 for a compact, spherical cluster. However,
Monte Carlo simulations of the full S&B model have
revealed that the protein-DNA clusters are not com-
pact [29], but rather have a surface that scales almost
linearly in m, close to the behavior of the simplified LC
012345JS10-1100101102103<n>m=50m=100m=200m=400012345JS10-310-210-1<n>/m(a)~exp(-JS)102m100101102103104105<n>JS=1JS=2JS=3JS=4102103m101102<n> eJS(b)~m0100200300400500s00.20.40.60.8ploop(s)JS=1JS=2JS=3JS=4(c)8
model presented here. The non-compact nature of the
protein-DNA cluster is perhaps not surprising because
each protein can form only one bridging bond.
A closely related statistic is the average accumulated
loop length (cid:104)(cid:96)(cid:105). From the LC partition function, we no-
tice that the loop length is completely decoupled from
the coupling constant JS and depends only on the up-
per cutoff (cid:96)max. Therefore, the cumulated average loop
length becomes:
(cid:104)(cid:96)(cid:105) =
(cid:96)max(cid:29)(cid:96)0
dν − 1
2 − dν
(cid:96)2−dν
max
(cid:96)1−dν
0
(cid:104)n(cid:105),
(13)
where the factor in front of (cid:104)n(cid:105) represents the average
length per loop. This prefactor induces a small algebraic
dependence on (cid:96)max, in contrast to (cid:104)n(cid:105) which depends
only on the lower cutoff (cid:96)0.
The loop statistics of protein-DNA clusters are not eas-
ily accessible in experiments. Instead, the most relevant
results for which this model can provide insight come
from ChIP-Seq experiments. These experiments yield
data for the enrichment of bound ParB as a function of
genomic position on the DNA, providing a measure of the
average protein binding profile of ParB on DNA [24, 25].
In the LC model, the ParB density profile along DNA
can be calculated from:
exp(−nJS) (14)
d(cid:96)n(cid:96)−dν
n Pn (s,{(cid:96)i})
(cid:96)0
m−1(cid:88)
n=0
PLC(s) =
1
ZLC
(cid:90) ∞
(cid:96)0
(m − n − 1)!n!
(m − 1)!
(cid:90) ∞
d(cid:96)1(cid:96)−dν
1
...
FIG. 5: Binding profiles of ParB from Eq. (14) plotted ver-
sus the genomic distance s to parS for (a) m = 100, (b) 200,
and (c) 400. In Eq. (14), the loop size integrals were calcu-
lated with a lower cutoff (cid:96)0 = 10 and an upper cutoff of 10(cid:96)0;
summations were truncated at n = 15. The dark grey circles
in panel (c) show experimental ChIP-Seq ParB enrichment
data from the F-plasmid of E. coli extracted from [25]. The
inset in panel (a) shows the binding profile of ParB versus ge-
nomic distance s to parS for JS = 1, ν = 0.588 (self-avoiding
polymer). The results in this inset were obtained by Monte
Carlo simulations of the LC model (see SI for details). The
data are plotted in log-log scale, we observe the power law
decay PLC ∼ s−dν as expected in the limit of low JS, where
the LC model becomes conceptually similar to the Stochastic
Binding model.
(cid:80)
(cid:80)n
where ZLC is given in Eq. (3). Here, Pn (s,{(cid:96)i}) repre-
sents the multiloop ParB binding profile with n loops of
length {(cid:96)i} = {(cid:96)1, ..., (cid:96)n}. For simplicity, we approximate
this multiloop profile by the analytical 1-loop conditional
probability, P1(s, (cid:96)loop(s)), with the loop length equal to
the accumulated loop length, i.e. (cid:96) →
i (cid:96)i, weighted by
the loop probability ploop(s,{(cid:96)i}) ≈
i=1 (cid:96)iρ(s, m, (cid:96)i, (cid:96)).
In the expression for the loop probability, ρ(s, m, (cid:96)i, (cid:96)) is
defined as the contribution to the loop density of a loop
of length (cid:96)i in a cluster of m proteins with a total ac-
cumulated loop length (cid:96), and we neglected correlations
between contributions from different loops. Furthermore,
we approximate ρ(s, m, (cid:96)i, (cid:96)) by using a generalization of
the 1-loop expression in Eq. (10),
ρ(s, m, (cid:96)i, (cid:96)) ≈
i + (cid:96)i(m + (cid:96) − 2(cid:96)i)] + Θ(2(cid:96)i − ((cid:96) + m))(cid:0) m+(cid:96)
min(s, (cid:96)i, m + (cid:96) − s)
2
Θ(m + (cid:96) − 2(cid:96)i) [(cid:96)2
(cid:1)2 .
(15)
In the analysis above, we aimed to capture the effects
of multiple loops in a simple way by assuming statisti-
0100200300s00.20.40.60.81PLC(s)100101102103s10-510-410-310-210-1100PLC(s)(a)PLC ~ s-dν0100200300400500600s00.20.40.60.81PLC(s)JS = 1JS = 2JS = 3JS = 4(b)cal independence of the loops, and by using the analyti-
cal 1-loop expressions to approximate the impact of loop
formation on the loop density and the ParB binding pro-
file of the protein-DNA complex. To test the validity of
these approximations, we performed MC simulations of
the complete LC model. We find that the numerically
obtained average loop probability is in reasonable agree-
ment with our approximate expression for the multiloop
density, as shown in Figure 4c. Thus, despite the simplic-
ity of our approach, the analytical model provided here
captures the essential features of looping in protein-DNA
clusters.
The protein binding profile PLC(s) around a parS site
is calculated by averaging the static binding profile for
different total loop numbers and loop lengths using the
Boltzmann factor (see Eq. (3)) from the Looping and
Clustering model as the appropriate weighting factor.
The resulting expression in Eq. (14) for the protein bind-
ing profile of a protein-DNA cluster is the central result
of this paper. We use this expression to compute bind-
ing profiles for the full Looping and Clustering model,
which are shown in Figure 5 as a function of the dis-
tance s to parS for m = 100, 200, and 400. By con-
struction, the site s = 0 corresponding to parS is always
occupied, and thus P (s = 0) = 1 for all values of the
spreading energy JS. This feature of the LC model cap-
tures the assumed strong affinity of ParB for a parS bind-
ing site. For JS = 4, the binding profile converges to a
triangular profile, implying a very tight cluster of pro-
teins on the DNA with almost no loops. The triangular
profile in this case results from all the distinct config-
urations in which this tight cluster can bind to DNA
such that one of the proteins in the cluster is bound to
parS, and therefore the probability drops linearly to 0
at s ≈ m. The same triangular binding profile was ob-
served for the S&B model in the strong coupling limit
JS → ∞ [29]. Interestingly, as JS becomes weaker, we
observe a faster decrease of the binding profile near parS
together with a broadening of the tail of the distribution
for distances far from parS. This behavior results from
the increase of the number of loops that extrude from
the ParB-DNA cluster with decreasing spreading bond
strength JS. The insertion of loops in the cluster allows
binding of ParB to occur at larger distances from parS.
Thus, the genomic range of the ParB binding profiles is
set by smax ≈ m + (cid:104)(cid:96)(cid:105), where the average cumulated
loop length (cid:104)(cid:96)(cid:105) is controlled by JS (see Eq. (13)) and m.
These results illustrate how the full average binding pro-
file is controlled by the spreading bond strength JS: the
weaker JS, the looser the protein-DNA cluster becomes,
which results in a much wider binding profile of proteins
around parS. In the limit JS → 0, the LC model quanti-
tatively reduces to the statistics of non-interacting loops,
as shown in the inset of Figure 5. In this case, the bind-
ing profiles exhibit asymptotic behaviour PLC(s) ∝ s−dν
for large s, as in the Stochastic Binding model [25]. In-
terestingly, we observe a weaker scaling PLC(s) with s
at intermediate genomic distances, which we attribute to
9
FIG. 6: Scaling function of the ParB binding profile for dif-
ferent total protein numbers m (same data as Figure 5). The
data for different total protein numbers m are plotted ver-
sus the dimensionless genomic distance s/m from parS (main
graph: JS = 3, inset: JS = 2).
the reduced loop density near parS (see Figure 4c).
To investigate how the functional shape of the binding
profile is determined by the total number of proteins in
the cluster, we plot the binding probability versus the
scaled variable s/m for m = 100, 200, and 400, as shown
in Figure 6. For fixed JS, the data approximately collapse
onto a single curve as a function of the scaled distance
s/m. This implies that the functional shape of the ParB
binding profile is largely determined by the spreading
bond strength JS, while the number of proteins in the
cluster determines the width of the profile.
V. DISCUSSION
The Looping and Clustering model introduced here
allows us to access the average binding profile of pro-
teins making up a large 3D protein-DNA complex.
In
our model, the formation of a coherent cluster of ParB
proteins is ensured by a combination of spreading and
bridging bonds between DNA bound proteins, which to-
gether can drive a condensation transition in which all
ParB proteins form a large protein-DNA complex local-
ized around a parS site [29]. We do not assume, however,
that this protein-DNA cluster is compact. Indeed, loops
of protein-free DNA may extend from the cluster, which
strongly influences the average spatial configuration of
proteins along the DNA. In the LC model, the forma-
tion of loops in the protein-DNA cluster is controlled by
the strength of spreading bonds, i.e. the bond between
proteins bound to nearest neighbor sites on the DNA.
Specifically, for every protein-free loop of DNA that ex-
tends from the cluster, a single spreading bond between
two proteins within the cluster must be broken. Thus, if
the spreading interaction energy, JS, is sufficiently small,
thermal fluctuations will enable the transient formation
and breaking of spreading bonds, thereby allowing mul-
tiple loops of DNA to emanate from the protein cluster
(See Figure 1).
Conceptually, the spreading bond interaction deter-
012s/m00.20.40.60.81PLC(s)00.511.5s/m00.20.40.60.81PLC(s)m = 100m = 200m = 400JS = 2JS = 3mines how "loose" the protein-DNA cluster is, which
directly impacts the ParB binding profiles. When JS
is large, loop formation is unlikely, resulting in a com-
pact protein-DNA cluster with a corresponding triangu-
lar protein binding profile centered around parS [29]. At
intermediate JS, the protein-DNA cluster becomes looser
with the formation of loops, resulting in a binding pro-
files that are more strongly peaked around parS but with
far-reaching tails.
Importantly, the LC model enables
us to establish a link between the Spreading & Bridging
model and the Stochastic Binding model [25]. The first
used a microscopic approach based on the types of in-
teractions between proteins on the DNA polymer, while
the second employed a more macroscopic approach based
on the polymer configurations around a dense sphere of
proteins. In the limit JS → 0, the LC model is consistent
with the Stochastic Binding model with a profile of the
form [25] given by P (s) ∝ s−dν(inset Figure 5a). Thus,
the LC model offers a description for a broad parameter
regime, connecting two limits investigated in preceding
studies [25, 29].
The Looping and Clustering model, which we intro-
duce to calculate the binding profile of ParB-like proteins
on the DNA, is a simple theoretical framework similar to
the Poland-Scheraga model for DNA melting [37, 38].
An important difference in the LC model with respect to
the homogeneous Poland-Scheraga model, is that trans-
lational symmetry is broken due to the presence of a
parS site at which a protein is bound with a high affinity.
Thus, the protein-DNA cluster can adopt a wide range of
configurations as long as one of the proteins is bound to
the parS site. As a result, loops are effectively excluded
in the vicinity of parS. The central new result of this work
is a simple way of computing the protein binding profiles
around such a parS site in terms of molecular interactions
parameters. We show that the binding profiles predicted
10
by this model are sensitive to both the expression level
of proteins and the spreading interaction strength, which
directly controls the formation of loops in the protein-
DNA cluster. The LC model predicts a profile in good
quantitative agreement with binding profiles measured
with ChIP-Seq on the F-plasmid of E. coli, as shown in
Fig. 5c. Importantly, from this analysis we extract the
spreading interaction strength JS ≈ 1kBT and the num-
ber of proteins in the cluster m ≈ 400.
Our results also have implications for experiments that
employ fluorescent labelling of DNA loci by exogenous
ParBs [26, 27]. Indeed, our model can be used to inves-
tigate how the protein interaction strengths determine
the 3D structure and mobility of the ParB-DNA cluster,
as well as the tendency of multiple ParB foci to adhere
to each other. This model thus provides an insightful
quantitative tool that could be employed to analyze and
interpret ChIP-Seq and fluorescence data of ParB-like
proteins on chromosomes and plasmids.
Acknowledgments
This project was supported by the German Excellence
Initiative via the program NanoSystems Initiative Mu-
nich (NIM) (C.P.B.), the Deutsche Forschungsgemein-
schaft (DFG) Grant TRR174 (C.P.B), and the National
Science Foundation Grant PHY-1305525 (N.S.W.). We
also thank J.-Y. Bouet for helpful comments on the
manuscript. The authors acknowledge financial support
from the Agence Nationale de la Recherche (IBM project
ANR-14-CE09-0025-01) and from the CNRS D´efi In-
phyniti (Projet Structurant 2015-2016). This work is also
part of the program "Investissements d'Avenir" ANR-
10-LABX-0020 and Labex NUMEV (AAP 2013-2-005,
2015-2-055, 2016-1-024).
[1] Dekker, J., Marti-Renom, M. A., & Mirny, L. A.
(2013). Exploring the three-dimensional organization of
genomes:
interpreting chromatin interaction data. Na-
ture Reviews Genetics, 14(6), 390-403.
[2] Scolari, V. F., & Lagomarsino, M. C. (2015). Combined
collapse by bridging and self-adhesion in a prototypical
polymer model inspired by the bacterial nucleoid. Soft
matter, 11(9), 1677-1687.
[3] Dame, R. T., Tark-Dame, M., & Schiessel, H. (2011).
A physical approach to segregation and folding of the
Caulobacter crescentus genome. Molecular microbiology,
82(6), 1311-1315.
[4] Jun, S., & Mulder, B. (2006). Entropy-driven spatial
organization of highly confined polymers:
lessons for
the bacterial chromosome. Proceedings of the National
Academy of Sciences, 103(33), 12388-12393.
[5] Emanuel, M., Radja, N. H., Henriksson, A., & Schies-
sel, H. (2009). The physics behind the larger scale or-
ganization of DNA in eukaryotes. Physical biology, 6(2),
025008.
[6] Marenduzzo, D., Micheletti, C., & Cook, P. R. (2006).
Entropy-driven genome organization. Biophysical jour-
nal, 90(10), 3712-3721.
[7] Mirny, L. A. (2011). The fractal globule as a model of
chromatin architecture in the cell. Chromosome research,
19(1), 37-51.
[8] Umbarger, M. A., Toro, E., Wright, M. A., Porreca, G. J.,
Bau, D., Hong, S. H., & Shapiro, L. (2011). The three-
dimensional architecture of a bacterial genome and its
alteration by genetic perturbation. Molecular cell, 44(2),
252-264.
[9] Le, T. B., Imakaev, M. V., Mirny, L. A., & Laub, M.
T. (2013). High-resolution mapping of the spatial orga-
nization of a bacterial chromosome. Science, 342(6159),
731-734.
[10] Viollier, P. H., Thanbichler, M., McGrath, P. T., West,
L., Meewan, M., McAdams, H. H., & Shapiro, L. (2004).
Rapid and sequential movement of individual chromoso-
mal loci to specific subcellular locations during bacterial
DNA replication. Proceedings of the National Academy
of Sciences of the United States of America, 101(25),
9257-9262.
[11] Lagomarsino, M. C., Esp´eli, O., & Junier, I. (2015). From
structure to function of bacterial chromosomes: Evo-
lutionary perspectives and ideas for new experiments.
FEBS letters, 589(20PartA), 2996-3004.
[12] Wang, X., Llopis, P. M., & Rudner, D. Z. (2013). Organi-
zation and segregation of bacterial chromosomes. Nature
Reviews Genetics, 14(3), 191-203.
[13] Dillon, S. C., & Dorman, C. J. (2010). Bacterial nucleoid-
associated proteins, nucleoid structure and gene expres-
sion. Nature Reviews Microbiology, 8(3), 185-195.
[14] Dame, R. T. (2005). The role of nucleoid?associated pro-
teins in the organization and compaction of bacterial
chromatin. Molecular microbiology, 56(4), 858-870.
[15] Mohl, D. A., & Gober, J. W. (1997). Cell cycle-dependent
polar localization of chromosome partitioning proteins in
Caulobacter crescentus. Cell, 88(5), 675-684.
[16] Lim, H. C., Surovtsev, I. V., Beltran, B. G., Huang, F.,
Bewersdorf, J., & Jacobs-Wagner, C. (2014). Evidence
for a DNA-relay mechanism in ParABS-mediated chro-
mosome segregation. Elife, 3, e02758.
[17] Banigan, E. J., Gelbart, M. A., Gitai, Z., Wingreen,
N. S., & Liu, A. J. (2011). Filament depolymerization
can explain chromosome pulling during bacterial mito-
sis. PLoS Comput Biol, 7(9), e1002145.
[18] Ptacin, J. L., Lee, S. F., Garner, E. C., Toro, E., Eckart,
M., Comolli, L. R., & Shapiro, L. (2010). A spindle-like
apparatus guides bacterial chromosome segregation. Na-
ture cell biology, 12(8), 791-798.
[19] Le Gall, A., Cattoni, D. I., Guilhas, B., Mathieu-
Demazi`ere, C., Oudjedi, L., Fiche, J. B., & Nollmann,
M. (2016). Bacterial partition complexes segregate within
the volume of the nucleoid. Nature Communications, 7.
[20] Walter, J. C., Dorignac, J., Lorman, V., Rech, J., Bouet,
J. Y., Nollmann, M., Palmeri, J., Parmeggiani A. & Ge-
niet, F. (2017). Surfing on protein waves: proteophoresis
as a mechanism for bacterial genome partitioning. Phys-
ical Review Letters 119(2), 028101.
[21] Vecchiarelli, A. G., Neuman, K. C., & Mizuuchi, K.
(2014). A propagating ATPase gradient drives transport
of surface-confined cellular cargo. Proceedings of the Na-
tional Academy of Sciences, 111(13), 4880-4885.
[22] Jindal, L., & Emberly, E. (2015). Operational principles
for the dynamics of the in vitro ParA-ParB system. PLOS
Comput Biol, 11(12), e1004651.
[23] Surovtsev, I. V., Campos, M., & Jacobs-Wagner, C.
(2016). DNA-relay mechanism is sufficient to explain
ParA-dependent intracellular transport and patterning of
single and multiple cargos. Proceedings of the National
Academy of Sciences, 113(46), E7268-E7276.
[24] Breier, A. M., & Grossman, A. D. (2007). Whole?genome
analysis of the chromosome partitioning and sporulation
protein Spo0J (ParB) reveals spreading and origin?distal
sites on the Bacillus subtilis chromosome. Molecular mi-
crobiology, 64(3), 703-718.
[25] Sanchez, A., Cattoni, D. I., Walter, J. C., Rech, J.,
Parmeggiani, A., Nollmann, M., & Bouet, J. Y. (2015).
Stochastic self-assembly of ParB proteins builds the bac-
terial DNA segregation apparatus. Cell systems, 1(2),
163-173.
[26] Chen, B., Guan, J., & Huang, B. (2016). Imaging specific
genomic DNA in living cells. Annual review of biophysics,
45, 1-23.
11
[27] Saad, H., Gallardo, F., Dalvai, M., Tanguy-le-Gac, N.,
Lane, D.,& Bystricky, K. (2014). DNA dynamics during
early double-strand break processing revealed by non-
intrusive imaging of living cells. PLoS Genet, 10(3),
e1004187.
[28] Rodionov, O., Lobocka, M., & Yarmolinsky, M. (1999).
Silencing of genes flanking the P1 plasmid centromere.
Science, 283(5401), 546-549.
[29] Broedersz, C. P., Wang, X., Meir, Y., Loparo, J. J., Rud-
ner, D. Z., & Wingreen, N. S. (2014). Condensation and
localization of the partitioning protein ParB on the bac-
terial chromosome. Proceedings of the National Academy
of Sciences, 111(24), 8809-8814.
[30] Graham, T. G., Wang, X., Song, D., Etson, C. M., van
Oijen, A. M., Rudner, D. Z., & Loparo, J. J. (2014). ParB
spreading requires DNA bridging. Genes & development,
28(11), 1228-1238.
[31] Taylor, J. A., Pastrana, C. L., Butterer, A., Pernstich,
C., Gwynn, E. J., Sobott, F., ... & Dillingham, M. S.
(2015). Specific and non-specific interactions of ParB
with DNA: implications for chromosome segregation. Nu-
cleic acids research, 43(2), 719-731.
[32] Gennes, P. G. D. (1979). Scaling concepts in polymer
physics.
[33] Leonard, T. A., Butler, P. J., & Lowe, J. (2005). Bacterial
chromosome segregation: structure and DNA binding of
the Soj dimer?a conserved biological switch. The EMBO
journal, 24(2), 270-282.
[34] Fisher, G. L., Pastrana, C. L., Higman, V. A., Koh, A.,
Taylor, J. A., Butterer, A., & Moreno-Herrero, F. (2017).
The C-Terminal Domain Of ParB Is Critical For Dy-
namic DNA Binding And Bridging Interactions Which
Condense The Bacterial Centromere. bioRxiv, 122986.
[35] Hanke, A., & Metzler, R. (2003). Entropy loss in long-
distance DNA looping. Biophysical journal, 85(1), 167-
173.
[36] Bouet, J. Y., & Lane, D. (2009). Molecular basis of the
supercoil deficit induced by the mini-F plasmid partition
complex. Journal of Biological Chemistry, 284(1), 165-
173.
[37] Poland, D., & Scheraga, H. A. (1970). Theory of helix-
coil transitions in biopolymers.
[38] Everaers, R., Kumar, S., & Simm, C. (2007). Unified
description of poly- and oligonucleotide DNA melting:
Nearest-neighbor, Poland-Sheraga, and lattice models.
Physical Review E, 75(4), 041918.
[39] Although this reasoning is not strictly true for self-
avoiding polymers, it does hold if we adopt the usual ap-
proximation used in the Poland-Scheraga model for DNA
melting that self-avoidance acts only within individual
loops
Appendix A: Monte Carlo simulations and
numerical integration procedures
1. Monte Carlo procedure
Using the partition function, we can formulate an effec-
tive 1D Hamiltonian for the LC model, which explicitly
accounts for the balance between spreading bonds and
loop entropy:
12
2. Numerical integration
To evaluate the binding profile PLC(s), we proceeded as
follows. We carried out the evaluation of the simplified
expression in Eq. (14) using numerical summation and
integration. We truncated the summation at n = 15,
instead of going up to m− 1, based on the corresponding
average number of loops of Fig. 5. Finally, we introduced
an upper cutoff for the loop-length, (cid:96)max = 100, instead
of going up to infinity. We confirmed that shape of the
binding profiles does not change significantly for higher
values of (cid:96)max = 100.
The numerical evaluation of the multidimensional in-
tegrals in Eq. (14) have been performed with an accuracy
and precision of respectively 2 and 3 effective digits in the
final results. We have carried out convergence tests of the
curve shapes in order to assess our parameters choice and
rule out numerical instabilities. All computations have
been performed by routines written in the Wolfram Lan-
guage and executed by the Mathematica software suite
(version 10 and 11).
Appendix B: Formal connection between the LC
model and a Lattice Gas with renormalized coupling
For m and n (cid:29) 1 (thermodynamic limit), we can for-
mulate a saddle point approximation to evaluate the par-
tition function and (cid:104)n(cid:105), by approximating the entropic
(factorial) term in Z (Eq.(11)) using the standard en-
tropy of mixing for placing n loops on m−1 possible sites.
This approach gives physical insight into how the loop
entropy contributes to a renormalized protein-protein in-
teraction and how the competition between this renor-
malized interaction and the entropy of mixing controls
(cid:104)n(cid:105). Taking the thermodynamic limit leads to a parti-
tion function:
(cid:90) ∞
0
Z(cid:48) =
dρ(cid:96) exp [−(m − 1)Feff (ρ(cid:96))] ,
(B1)
where ρ(cid:96) = n/(m − 1) is the concentration of loops (0 ≤
ρ(cid:96) ≤ 1) and
Feff (ρ(cid:96)) = ρ(cid:96)J(cid:48)S + {[1 − ρ(cid:96)] ln[1 − ρ(cid:96)] + ρ(cid:96) ln[ρ(cid:96)]} (B2)
an effective free energy where J(cid:48)S = JS +ln α0 is a loop ac-
tivation energy renormalized by the cost in loop entropy
with α0 = (cid:96)dν−1
(dν − 1). In the limit m → ∞, the ap-
proximate partition function Z(cid:48) becomes exact and can
be evaluated exactly in the saddle point approximation
by minimizing Feff . The solution, ρSP, to the saddle point
equation, dFeff (ρ(cid:96))/dρ(cid:96) = 0, is
0
ρSP =
1
1 + eJ(cid:48)
S
.
(B3)
The entropic contribution to Feff (second term) vanishes
at ρ(cid:96) = 0 and 1, and reaches a minimum at ρ(cid:96) = 1/2,
FIG. 7: The binding profile obtained with the analytic ap-
proach (symbols) are compared to MC simulations (dashed
lines) for m = 100, (cid:96)0 = 10, and JS = 1, 2, 3 and 4.
L−1(cid:88)
i=1
HLC = −JS
n(cid:88)
i=1
φiφi+1 + dν
ln((cid:96)i + (cid:96)0) .
(A1)
This effective Hamiltonian is useful to perform Monte
Carlo simulations of the model as a benchmark for the
approximations performed in the analytical approach
(see Fig.7). The proteins are modelled as particles that
bind/unbind onto sites of a one-dimensional lattice with
free boundary conditions. The lattice size L = 4000 is
chosen to prevent finite size effect for the range of pro-
teins considered. Note that, in these MC simulations, the
total size of the loops is limited to L − m.
Metropolis rules:
The simulations are performed with the standard
1. Propose a move of a particle randomly chosen to
a random empty site of the lattice (conserved or-
der parameter). A MC iteration step consists of m
attempts of move.
2. Calculate the difference of energy ∆H = Hf − Hi
between final and initial configurations.
3. If ∆H < 0, the move is accepted with proba-
bility 1, otherwise it is accepted with probability
exp(−β∆H).
The system is set initially with all particles in a single
cluster (JS = ∞), and then thermalized to the actual JS
of the simulations ranging from 1 to 4 (see Fig. 7). The
sampling starts after thermalization of the system (40000
MC iterations). A sampling of the systems configuration
is performed every 100 MC iterations. All MC averages
have been performed over 107 configurations, ν = 0.588,
L = 4000 and (cid:96)0 = 10. The numerical results of this
Monte Carlo simulation are in good agreement with our
approximate analytic results, as shown in Fig. 7.
0100200300s00.20.40.60.81PLC(s)JS=1JS=2JS=3JS=4(a)m = 100which is the exact result for ρSP at vanishing renormal-
ized loop activation energy J(cid:48)S because the entropy of
mixing is then maximized. For J(cid:48)S > 0, ρSP decreases
(cid:48)
from 1/2 to vanish in the limit J(cid:48)S → ∞ as ρSP → e−J
S .
In this limit only the no and one loop states contribute
and the asymptotic behavior can be simply obtained by a
series expansion of the partition sum and the correspond-
ing expression for (cid:104)n(cid:105). The saddle point result leads to
nSP =
m − 1
1 + eJ(cid:48)
S
,
(B4)
13
which turns out to be the exact result, thanks to compen-
sating errors, for (cid:104)n(cid:105) for finite m (which can be obtained
by differentiating the exact Z with respect to JS, see
Eq.(12)). For example, for (cid:96)0 = 10, d = 3, and ν = 0.588,
α0 = 4.437 and ln(α0) = 1.49, which is not negligible if
JS = O(1).
|
1506.08168 | 2 | 1506 | 2015-11-24T18:57:28 | Shaping the Growth Behaviour of Biofilms Initiated from Bacterial Aggregates | [
"physics.bio-ph",
"q-bio.CB"
] | Bacterial biofilms are usually assumed to originate from individual cells deposited on a surface. However, many biofilm-forming bacteria tend to aggregate in the planktonic phase so that it is possible that many natural and infectious biofilms originate wholly or partially from pre-formed cell aggregates. Here, we use agent-based computer simulations to investigate the role of pre-formed aggregates in biofilm development. Focusing on the initial shape the aggregate forms on the surface, we find that the degree of spreading of an aggregate on a surface can play an important role in determining its eventual fate during biofilm development. Specifically, initially spread aggregates perform better when competition with surrounding unaggregated bacterial cells is low, while initially rounded aggregates perform better when competition with surrounding unaggregated cells is high. These contrasting outcomes are governed by a trade-off between aggregate surface area and height. Our results provide new insight into biofilm formation and development, and reveal new factors that may be at play in the social evolution of biofilm communities. | physics.bio-ph | physics | Shaping the Growth Behaviour of Biofilms Initiated from Bacterial Aggregates
Gavin Melaugh,1 Jaime Hutchison,2 Kasper Nørskov Kragh,3 Yasuhiko Irie,4, 5 Aled Roberts,4
Thomas Bjarnsholt,3, 6 Stephen P. Diggle,4 Vernita D. Gordon,2 and Rosalind J. Allen1
1School of Physics and Astronomy, University of Edinburgh,
James Clerk Maxwell Building, Peter Guthrie Tait Road, Edinburgh, EH9 3FD
2Center for Nonlinear Dynamics and Department of Physics,
The University of Texas at Austin, Austin, Texas 78712-1199
3Department of International Health, Immunology and Microbiology,
Faculty Of Health Sciences, University of Copenhagen, DK-2200 Copenhagen
4School of Life Sciences, Centre for Biomolecular Sciences,
University of Nottingham, University Park, Nottingham NG7 2RD
5Department of Biology & Biochemistry, University of Bath, Claverton Down, Bath BA2 7AY
6Department for Clinical Microbiology, University of Copenhagen, DK-2100 Copenhagen
Bacterial biofilms are usually assumed to originate from individual cells deposited on a surface.
However, many biofilm-forming bacteria tend to aggregate in the planktonic phase so that it is
possible that many natural and infectious biofilms originate wholly or partially from pre-formed cell
aggregates. Here, we use agent-based computer simulations to investigate the role of pre-formed
aggregates in biofilm development. Focusing on the initial shape the aggregate forms on the surface,
we find that the degree of spreading of an aggregate on a surface can play an important role in
determining its eventual fate during biofilm development. Specifically, initially spread aggregates
perform better when competition with surrounding unaggregated bacterial cells is low, while initially
rounded aggregates perform better when competition with surrounding unaggregated cells is high.
These contrasting outcomes are governed by a trade-off between aggregate surface area and height.
Our results provide new insight into biofilm formation and development, and reveal new factors that
may be at play in the social evolution of biofilm communities.
INTRODUCTION
Surface-attached communities known as biofilms are
believed to be the predominant mode of existence for bac-
teria in many environmental settings [1]. Understanding
how biofilms establish and grow is also clinically impor-
tant given their ubiquity in medical implant infections [2],
chronic wounds [3], and in the respiratory tracts of cystic
fibrosis patients [4]. In the clinical context, biofilm com-
munities often show enhanced virulence [5], resistance to
antibiotics [6], and resistance to the host immune system
[7]. These features may be associated with the spatial
structure of the biofilm, which not only affects material
transport, e.g., penetration of nutrients/antibiotics, but
is also associated with differences in metabolism and gene
expression among cells within the community [8, 9].
In the canonical picture of biofilm development, indi-
vidual cells land on a surface, attach and proliferate to
form first micro-colonies and later 3-dimensional struc-
tures [10]. However, bacteria are also known to form
dense aggregated clumps when they are grown in liquid
(planktonic phase) [11 -- 13]. Moreover, cells often disperse
from existing biofilms as clumps of aggregated cells. Thus
it is very likely that when a biofilm forms, some cells may
arrive on the surface already in an aggregated state. In
support of this view, evidence exists for the seeding of in-
fections by pathogenic bacteria already in an aggregated
state [14, 15], and bacterial aggregates are abundant in
cystic fibrosis [4, 5] and tuberculosis [16] infections.
Having arrived on the surface, e.g., a plant leaf [17], a
surgical implant [2] or an industrial component [18], it is
to be expected that cells within a bacterial aggregate will
have to compete during biofilm development, both with
other aggregates and with initially non-aggregated cells,
to which they may or may not be genetically related.
We take a first step towards understanding the role of
pre-formed aggregates in biofilm development by investi-
gating this competitive process, using agent-based simu-
lations. Such simulations, in which the spatial structure
of a biofilm emerges from local interactions between in-
dividual cells, have become a staple tool for investigating
biofilm structure and dynamics [19 -- 21], as well as so-
cial evolutionary aspects of biofilm development [22, 23].
Using this approach, we determine how a pre-existing
aggregate of bacteria impacts the spatial structure of a
biofilm, both in the presence and absence of competing
unaggregated bacterial cells.
Our main focus here is on the role of the initial shape
of the aggregate. It is well known that bacterial inter-
actions with a surface depend on features such as extra-
cellular polymeric substances (EPS), presence of cell sur-
face appendages (such as pili), and cell surface charge,
which are species- and strain-dependent [24]. Moreover,
soft-matter science has established that the nature of
material-surface interactions can drastically affect the
shape of fluid or semi-fluid droplets on surfaces [25]. It
is therefore reasonable to suppose that in some circum-
stances, bacterial aggregates will spread out in contact
with a surface, while in other scenarios, aggregates will
adopt a more compact configuration. Here we investi-
gate the biological consequences of aggregate shape in
the seeding of biofilm growth.
arXiv:1506.08168v2 [physics.bio-ph] 24 Nov 2015
2
gated cells has been suggested to be a stepping stone in
the evolution of multicellularity [26, 27]
Our study highlights the effects of nutrient gradients
and bacterial aggregate shape on long-term biofilm devel-
opment. Our work reveals that these factors alone can
produce a trade-off between nutrient access and compe-
tition, with the balance between these factors depending
sensitively on aggregate shape. While the link between
biofilm spatial structure and nutrient access has been
highlighted in many other studies [8, 23, 28 -- 30], our work
is the first to focus on the role of preformed aggregates
in this context. Our study should help to decipher the
role of pre-formed aggregates in biofilm infections. More
generally, our findings emphasise the need to consider
pre-formed aggregates in our current understanding of
biofilm development.
METHODS
In this study, we use agent-based computer simula-
tions to model the growth of a biofilm on a surface, start-
ing from initial configurations of bacterial cells like those
shown in Figure 1(a). In our simulations, an initial ag-
gregate of cells (shown in green in Figure 1(a)), adopting
a particular shape, seeds an inert surface, and may com-
pete with surrounding unaggregated cells (red in Figure
1(a)). Note that the red and green bacterial cells differ
only in the manner in which they are initially arranged
on the surface. At the start of our simulations, the "red"
bacteria are distributed at random across those parts of
the surface not occupied by the aggregate (see Sections
S1 to S5). To vary the extent of competition between the
aggregated and unaggregated cells, we varied the initial
cell density (number of cells per unit length of surface) of
the unaggregated "red" cells (see Sections S1 to S5). As
a control, we also ran simulations in which the aggregate
grew in the absence of the unaggregated cells.
The focus of this work is on the shape of the initial
cell aggregate. Figure 1(a) illustrates three different sce-
narios,
in which the cell aggregate adopts a compact
rounded shape (top), spreads out on the surface (bot-
tom), or adopts an intermediate shape (middle). In each
scenario, the dimensions of the aggregate are adjusted so
that the number of cells within the aggregate remains the
same (see Sections S1 and S2).
Aggregate Shape Characteristation
To characterise quantitatively the aggregate shapes in
the different scenarios shown in Figure 1(a), we define
the "aggregate-surface angle" θ, which is the angle that
the initial aggregate makes with the flat surface. A small
value of θ (θ → 0◦) describes an initial aggregate config-
uration that is spread on the surface, whereas a large
value of θ (θ → 180◦) describes an aggregate that is
rounded. Given that the total number of cells in the
FIG. 1. Our simulation set-up. (a) Schematic representation
of bacterial aggregates (green) which are initially spread on
a surface to varying extents. The schematic also shows sur-
rounding, competing, unaggregated cells (red). Top- Rounded
aggregate, θ = 180o; Middle- Semi-spread aggregate, θ = 90◦;
Bottom- Spread aggregate with θ → 0◦. Note that the size
of the aggregates (in terms of number of bacteria) is approxi-
mately equal. (b) Growth of the spread (θ = 5o), semi-spread
(θ = 90◦), and rounded aggregate (θ = 180◦) populations over
the course of our simulations in the absence (ρ = 0 cell µm−1)
and presence of competition (ρ = 0.5 cell µm−1). For clarity
the error bars, representing the standard deviations, are only
shown for the final data points. The standard deviations at
these points are maximal.
Simulating the development of biofilms initiated from
initially spread or rounded aggregates, we find that the
initial configuration of a bacterial aggregate on a sur-
face is crucial in determining its eventual fate within
the biofilm.
In the absence of competitor cells on the
surface, aggregates that maximise the extent to which
they initially spread on the surface perform better than
rounded ones because their initial access to nutrients (in
the surrounding media) is greater. However when faced
with strong competition from neighbouring unaggregated
cells, initially rounded aggregates perform better at long
times, despite the fact that the rounded aggregate shape
has a smaller surface area and hence a reduced expo-
sure to nutrients. Importantly, we show that in an ini-
tially rounded aggregate, cells at the top of the aggregate
proliferate at the expense of cells in the aggregate cen-
tre. This has interesting possible consequences for social
evolution given that cooperation within clumps of aggre-
Spread, no competition
Spread, high competition
Semi-spread, no competition
Semi-spread, high competition
Rounded, no competition
Rounded, no competition
0
50 100 150 200 250 300 350 400 450 500
time [h]
16000
14000
12000
10000
8000
6000
4000
2000
0
population of aggregate cells, N
initial aggregate is fixed, θ also encapsulates an interplay
between the initial surface coverage of the aggregate and
its initial height; with increasing θ, the surface coverage
decreases whereas height increases.
In soft matter sci-
ence, an analogous parameter is often used to describe
wetting interactions between liquid droplets and surfaces
[25]. A similar approach has recently been applied to the
surface-spreading behaviour of eukaryotic cell aggregates
[31]. From a phenotypic perspective, θ is related to the
nature of the interactions between cells in the aggregate
and between cells and the surface, and thus could be tun-
able by biological regulatory processes, or by evolution.
We performed simulations for a range of θ values between
5◦ and 180◦ (Figure 1(a)).
The aggregate configurations that we used to initiate
our simulations were generated by "transplantation" of
circular segments from simulation snapshots of pre-grown
biofilms (see Sections S1 and S2). This procedure proved
preferable to other initialisation methods as it ensures no
overlap between individual bacteria and enables the gen-
eration of different aggregate shapes of the same number
density (∼100 cells per unit area). By varying the radius
of the circular segment we are able to ensure that each
aggregate contained ∼100 cells that were initially spread
on the surface to different extents. To ensure statistical
accuracy, four different configurations were generated for
each aggregate shape, defined by its value of θ, and for
each of these configurations, five simulations were per-
formed using different random number seeds. Changing
the random number seed affects the order in which in-
dividual bacteria grow and divide, and also changes the
locations of the unaggregated cells on the surface sur-
rounding the aggregate. A total of twenty simulations
were therefore performed for each value of θ, enabling us
to sample both variation in the configuration and the or-
dering of cell updates (Section S5). Increasing the num-
ber of repeated simulations did not affect our results.
In common with many other biofilm simulation studies
[19, 22, 23, 32], our simulations were performed in two
dimensions for the purposes of computational efficiency.
We have verified, however, that our key findings are re-
produced when we use 3D simulations (see Sections S3
and S10).
Simulation Implementation
We use the agent-based microbial simulation package
iDynoMiCs [33] to model biofilm growth, starting from
configurations such as those shown in Figure 2. In these
simulations, individual bacterial cells are represented as
spherical agents, which grow and proliferate conditional
on the local nutrient concentration, and "shove" each
other apart to relieve local stresses within the biofilm.
The order in which cells are selected to grow and divide
is random during each global time-step of the simulation,
as is the direction of cell division. In our simulations, the
initial distribution of surrounding competitor cells on the
3
surface is also random. The simulations use a spatial grid
to track the local nutrient concentration field. Nutrient is
assumed to diffuse towards the biofilm from above, with
the concentration being fixed to a bulk value in a layer
far from the biofilm. Within the biofilm itself, nutrient
diffusion is hindered relative to the region outside the
biofilm. Nutrient consumption by the bacterial cells leads
to local gradients, which can have a strong impact on the
structural features of the growing biofilm [8, 23, 28 -- 30].
Periodic boundary conditions are imposed on both the
nutrient concentration field and the particle coordinates
in the horizontal direction.
From a mathematical perspective, nutrient is repre-
sented as a concentration field, the dynamics of which
are governed by the the reaction-diffusion equation
∂S(x)
∂t
= ∇ · (DS(x) · ∇S(x)) + rS(x),
(1)
where S(x) is the space (x)-dependent nutrient concen-
tration, DS(x) is the diffusion coefficient of the nutrient,
and rS(x) is the consumption rate of the nutrient by the
bacteria. The rate of nutrient consumption, rS(x), is
related to the growth rate of the bacteria, dX/dt, via
rS(x) =
dS
dt
= −
1
Yx/s
dX
dt
,
(2)
where X(x) is the local biomass density, and Yx/s is a
yield coefficient that describes the amount of nutrient
required to produce one unit of biomass X.
The growth rate of each cell is governed by the well-
known Monod function
dX
dt
= µmax
S
kS + S
X,
(3)
where µmax is the maximum specific growth rate of the
bacteria, and kS is the concentration of nutrient, S, at
which the growth rate is half maximal. The growth pa-
rameters used in our simulations were taken from em-
pirical and simulation studies on Pseudomonas aerugi-
nosa, assuming glucose to be the rate-limiting nutrient
(see Table I). However our results are not sensitive to
the detailed parameter choice, or to the choice of rate-
limiting nutrient. Note that the growth rate parameters
YX/N , µmax, and kS are the same for both the aggregate
cells and the competitor cells (see Table I).
From a practical point of view, in idynoMiCs the nu-
trient concentration fields are assumed to be in pseudo
steady-state with respect to biomass growth and there-
fore the time dependence is removed from Equation 1
0 = ∇ · (DS(x) · ∇S(x)) + rS(x).
(4)
This equation is solved numerically for every global up-
date of the bacterial population. In our simulations, we
used a bulk nutrient concentration of 5.4 × 10−3 gL−1
(see Section S6) comparable with previous work [32 -- 34].
Each of our simulations was run for a total time of 480 h,
Yield coefficient for Monod equation (Equation 3) 0.44
Symbol Description
[S]bulk Bulk concentration of limiting nutrient
YX/N
µmax Maximum specific growth rate
KN
γ
DG
Lx
Ly
Ldbl
Half saturation concentration of nutrient
Density of biomass
Diffusivity of glucose in water
Dimension of system in horizontal direction
Dimension of system in vertical direction
Thickness of diffusion boundary layer
4
Notes/ref
Value
5.4 × 10−3 gL−1 Within range of values from [32 -- 34]
0.35 h−1
3 × 10−3
200 gL−1
5.8−5 m2day−1
1032 µm
1032 µm
80 µm
Within range of values from [32, 35, 36]
Within range of values from [35, 37 -- 39]
Within range of values from [20, 21, 33, 35 -- 38, 40]
[20, 32, 35]
[41]
Ensures aggregates do not interact periodically
Corresponds to the horizontal length
Within range of values from [20, 21, 42]
TABLE I. Input parameters for biofilm simulations.
in order to explore the long-term growth dynamics (see
results and Section S7).
Our complete parameter set is listed in Table I. Using
the parameters in Table I, our simulations produce spa-
tially structured biofilms 200-300 µm in height after a
simulation time of 480 h (see Figure 5, and Section S6).
For an aggregate in the presence of competing non-
aggregating cells, Figure 1(b) points to two strategies
for maximising progeny. For long-lived biofilms, progeny
can be maximised by the aggregate adopting a rounded
configuration, whereas if the biofilm is short-lived then
it may instead be optimal for the aggregate to spread on
the surface.
RESULTS
Initial Aggregate Shape Determines Growth
Dynamics
To assess the growth dynamics of pre-formed aggre-
gates in a biofilm, we tracked the number of progeny cells,
N , produced by aggregates of different shape (Figure
1(b)). We investigated three different aggregate shapes,
characterised by the angle θ (See Methods). The three
angles, θ = 5◦, 90◦, 180◦, describe pre-formed aggregates
that are initially arranged on the surface in either a
spread, intermediately spread, or rounded manner. To
investigate the effect of the surrounding unaggregated
competitor cells (red cells in Figure 1(a)), we varied the
density, ρ, of these cells between two extremes regimes of
competition reported in this study: ρ = 0 cell µm−1, no
competition; and ρ = 0.5 cell µm−1, high competition.
Not surprisingly, aggregates grow better in the absence
of surrounding cells on the surface regardless of their
initial shape, i.e., for ρ = 0 cell µm−1.
In this "non-
competitive regime", the initially spread aggregate pro-
duces more progeny than the more rounded aggregate.
This is evident in Figure 2, which shows representative
initial aggregate configurations and the structures of the
biofilms which they form after 480 hours.
Competition from unaggregated competitor cells on
the surfaces leads to more complex behaviour. Figure
1(b) shows that, in the presence of strong competition,
the spread aggregate produces more progeny over short
times than the rounded aggregate. However, over longer
times, the size of the population arising from the rounded
aggregate is larger.
Initial Aggregate Shape Affects Long-Term Biofilm
Structure
In our simulations the initial shape of the aggregate
influences the long-term structure of the biofilm. Fig-
ure 2 shows typical biofilm structures formed after 480
h of growth, starting from aggregates that were initially
spread on the surface (θ = 5◦, left), rounded (θ = 180◦,
right) or partially spread (θ = 90◦, centre), in the absence
of competition from surrounding cells. It is clear that the
spread aggregate covers much more of the surface during
growth than its more rounded counterparts.
In the presence of competition (Figure 3), we observe
a marked difference in the structure of the biofilms that
originate from spread aggregates (left panels) and from
rounded aggregates (right panels). For the spread aggre-
gate (a and c), the green section of biofilm that originates
from the aggregate is structurally indistinguishable to
that of the surrounding red biofilm that originated from
the competing, unaggregated cells. In contrast, for the
rounded aggregate (b and d), cells originating from the
aggregate form a distinct "clump", which is taller than
the surrounding biofilm. When the density of competing
(red) cells is high (Figure 3(d)), there is a cell-free gap
around the growing clump that appears to be a result of
nutrient depletion.
It is clear from the biofilm structures shown in Figures
2 and 3 that, even at very long times, the spatial structure
of a biofilm can be affected by the initial spatial configu-
ration of its founder cells (see also Section S6). While it
might seem remarkable that apparently small changes in
initial configuration can have dramatic effects on biofilm
structure even after many cell generations, this effect is
5
FIG. 2. Initial aggregate arrangement affects biofilm morphology. Simulation snapshots of three bacterial aggregates initially
arranged on the surface and the biofilms they form after 480 h: (a) Spread, 0 h. A zoomed in image is also shown to make the
shape of the aggregate easier to resolve; (b) Semi-spread, 0 h; (c) Rounded, 0 h; (d) Spread, 480 h; (e) Semi-spread, 480 h; (f)
Rounded, 480 h.
are effectively created by the presence of the initial ag-
gregates. The introduction of the rounded aggregate
amongst the lawn of unaggregated cells on the surface
at high competition leads to an instability in the biofilm
structure that propagates as the biofilm develops.
Nutrient gradients are important determinants of
aggregate fate
FIG. 3. Aggregate shape and neighbouring strain density
affect biofilm morphology. Simulation snapshots of biofilms
seeded from spread and rounded aggregates after 480 h growth
in the presence of a low and high density inoculum of the
competing strain: (a) θ = 5◦, ρ = 0.01 cell µm−1; (b) θ =
180◦, ρ = 0.01 cell µm−1; (c) θ = 5◦, ρ = 0.5 cell µm−1; (d)
θ = 180◦, ρ = 0.5 cell µm−1.
in fact well-known in a different context. For initially
flat biofilms, Dockery and Klapper showed theoretically
that small inhomogeneities in initial configuration may be
magnified into large "fingers" over the course of biofilm
development [30]. This phenomenon is shown as a finger-
ing instability and arises from the fact that an emerging
protrusion (or finger) is elevated above, and thus depletes
nutrients from the surrounding biofilm. This leads to
positive feedback, in which the enhanced growth of the
cells at the top of the instability is to the detriment of
the surrounding cells below [43].
While Dockery and Klapper assumed that struc-
tural inhomogeneities would arise spontaneously during
biofilm growth, in our simulations such inhomogeneities
FIG. 4. Cells on the outside of the aggregates grow faster
because they have greater access to nutrients. (a) Cell growth
rate (µ) distribution of the biofilm formed from the semi-
spread aggregate in the absence of competition after 4h. (b)
Corresponding nutrient concentration field, [S].
6
FIG. 5. Gradients in individual cell growth rates emerge in our simulated biofilms during growth. Cell growth rate distributions
for the spread and rounded aggregates at after 480 h of growth: (a) θ = 5◦, ρ = 0.0 cell µm−1; (b) θ = 5◦, ρ = 0.01 cell µm−1;
(c) θ = 5◦, ρ = 0.5 cell µm−1; (d) θ = 180◦, ρ = 0.0 cell µm−1; (e) θ = 180◦, ρ = 0.01 cell µm−1; (f) θ = 180◦, ρ = 0.5 cell
µm−1. These distribution correspond to the configurations in Figures 2 and 3. Note that the gradient in cell growth rate is so
large that a log scale is used for visualisation purposes. The green dashed lines represents an approximate boundary between
the aggregate cells and the surrounding competing strain.
It is well known that growth rate heterogeneities, re-
sulting from nutrient concentration gradients, emerge
during biofilm growth [44]. With this in mind, we tracked
the growth rates of individual cells as a function of their
position within the growing biofilm. Even in the very
early stages of biofilm growth, we see heterogeneity in
growth rates which emerge from (and influence) spatial
gradients in nutrient concentration. Figure 4 illustrates
this for a semi-spread aggregate (θ = 90o), after 4 h
of growth, in the absence of competition. As expected,
the cell growth rate is highly heterogeneous across the
biofilm, Figure 4(a), with cells on the outside growing
faster than those on the inside because they have better
access to nutrients (Figure 4(b)).
The growth rate heterogeneities shown in Figure 4(a)
are amplified in the later stages of biofilm growth. Figure
5 shows the spatial distribution of cell growth rates for
biofilms arising from spread and rounded aggregates after
480 h, in the presence and absence of competitor cells,
for the same simulations as shown in Figures 2 and 3.
In all cases we observe, as in previous work [45], two
distinct regions of growth activity within the develop-
ing biofilms: an outer layer of metabolically active cells
and an interior region of inactive cells. These distinct
regions arise because consumption by cells in the outer
layer deprives cells in the inner layer of nutrients [45].
We also observe a large gradient in individual cell growth
rate within the growing layer itself (note the logarithmic
scale in Figure 5). The dynamics of the metabolically
active layer determine the overall growth behaviour and
structure of the biofilm. In Section S9 we show that the
active layer of the rounded aggregate, unlike the spread
aggregate, continues to expand in the presence of compe-
tition; explaining why its total population becomes larger
than that of the spread at longer times in Figure 1(b),
and why it's structures tend to fan outwards at the top
(Figure 5(f)).
Although it is well documented that nutrient gradi-
ents arising during biofilm growth play an essential role
in biofilm formation [8, 23, 29], so far few studies have
investigated the effect of pre-formed bacterial clumps in
this process; in particular how the initial arrangement of
cells within a clump affects biofilm structure and develop-
ment. Figures 4 and 5 show that the initial arrangement
of cells on the surface can determine the shape and struc-
ture of a growing biofilm because small initial differences
in nutrient gradients become amplified as the biofilm de-
velops.
Competition for Nutrient Favours Rounded
Aggregates
Next we investigated how the fate of an aggregate, as
measured by the average number of progeny of one of
its cells, varies with aggregate shape. To this end, we
computed the number of progeny cells, N , arising from
the aggregate after a period of biofilm growth, relative
to the initial number of cells in the aggregate, N0, for a
range of aggregate shapes, determined by θ, at varying
7
FIG. 6. Average number of progeny, N/N0, of aggregates
defined by their surface-aggregate angle θ, the functional from
of which changes with increasing density of competitor cells:
(a) ρ = 0 µm cell−1; (b) aggregate-medium interface length,
s, as a function of θ; (c) ρ = 0.145 µm cell−1; ρ = 0.5 µm
cell−1. Vertical bars in represent the standard deviation from
20 data points.
levels of competition. For this analysis, we carried out
long simulations (480 h of biofilm development), so that
the ratio N/N0 reflects the long-time fate of the progeny
of cells within the aggregate (see Section S7).
Figure 6(a) shows N/N0 plotted as a function of the
aggregate-surface angle, θ, in the absence of competition
from surrounding unaggregated cells. It is clear that the
spread aggregate produces more progeny on average than
the rounded one.
In the previous section we saw that cells on the out-
side of the aggregate have more access to nutrients even
in the very early stages of biofilm growth. Thus, we might
hypothesise that the growth advantage of the spread ag-
gregate, in the absence of competition, is related to its
larger surface area. Indeed, Figure 6(b) shows that N/N0
correlates closely with the interfacial area (or arc length)
of the initial aggregate, s(θ). We therefore conclude that
the spread aggregate produces more progeny than the
rounded one because the former has a greater surface
area with the surrounding medium, providing greater ex-
posure to nutrient in the initial stages of growth. The
difference in initial structure between the spread and
rounded aggregates therefore translates into significant
differences in cell fate, even after many generations of
biofilm growth.
As the density of competition of unaggregated cells on
the surface increases, however, a very different scenario
emerges. Figures 6(c) and (d) show that cells in the
rounded aggregate (large θ) produce more progeny, on
average, than those in the spread aggregate. This is more
evident in Figure 7, which shows N/N0 for the spread
and rounded aggregates as a function of the density of
competitor cells. This effect is also observed at higher
nutrient concentration (see Section S8).
FIG. 7. Relative fitness as measured by N/N0 of rounded
aggregates increases with competition. Rounded aggregates
become favourable relative to spread aggregates with increas-
ing density of competitor cells.
Why does competition from surrounding unaggregated
cells favour rounded over spread aggregates? Close in-
spection of Figure 6 shows that, while the number of
progeny produced by the spread aggregate decreases with
increasing competition from surrounding cells (panels
(c), and (d)), the number of progeny produced by the
rounded aggregate remains rather constant.
This finding can be understood by investigating how
the fate of an individual cell within an aggregate depends
on its initial spatial location. To this end, we tracked the
number of progeny of each individual founder bacterium,
as a function of its initial position within an aggregate.
This constitutes a local, spatially-resolved version of the
"fitness measure" N/N0. Averaging our results over 20
repeated simulations allowed us to generate a map show-
ing the average number of progeny produced by indi-
vidual cells within an aggregate, for initially spread and
rounded aggregates, Figure 8.
Figure 8 shows that the initial position within an ag-
gregate indeed has a strong effect on cell fate.
In the
absence of competition from surrounding unaggregated
cells, the most successful cells in the spread aggregate
are those at the horizontal extreme edges; in the interior
region of the aggregate, cell fate is more uniform (Figure
8(a)). It seems likely that in this case, cells at the hori-
zontal edges have an advantage because their progeny can
expand in the horizontal direction, whereas the progeny
of cells in the interior of the aggregate must compete
with their neighbours within the aggregate for nutrients
and space. The proliferation of the cells at the edges of
the aggregate drives the lateral expansion of the growing
biofilm which we observe in Figures 2(a) and (d). In con-
trast, for the rounded aggregate (Figure 8(b)), cell fate
is overall more heterogeneous within the aggregate, with
the most successful cells being located around the out-
0 20 40 60 80 100 120 140 160 180
0 20 40 60 80 100 120 140 160 180
θ [
o
]
θ [
o
]
(a)
(d)
300
250
200
150
s [µm]
100
50
55
50
45
40
35
30
25
0
N / N
(b)
(c)
145
140
135
130
125
120
115
110
105
0
N / N
55
50
45
40
35
30
25
0
N / N
***
P = 0,0008
**
***
P = 0,0031
P < 0,0001
***
P < 0,0001
Spread
Rounded
80
60
40
20
0
N/N0
0.01
Density of Competitor Cells [cell um-1]
0.145
0.291
0.5
side surface of the aggregate. For the rounded aggregate,
it appears that height is a relevant factor as well as the
proximity to the aggregate surface.
Figure 8(c) shows that, in the presence of competition,
cells at the horizontal edges of the spread aggregate actu-
ally do less well than those in the centre. The decreased
fitness of these cells explains the inhibited lateral expan-
sion observed in Figures 3(a) and (c). For the rounded
aggregate, the most successful cells in the absence of com-
petition are those at the top of the aggregate.
In the
presence of competition (Figure 8(d)), these cells, which
are now highly localised at the top, are elevated above
the level of the competitor cells and therefore are little
affected by the increased competition for nutrients. The
"fitness" cost associated with its smaller surface inter-
face is compensated in the presence of competition by its
height, since its top cells remain unchallenged by com-
petitors with respect to nutrient access.
DISCUSSION
Given the tendency of many bacteria to aggregate, and
the frequent observation of aggregates in diverse environ-
mental situations [5, 46, 47], it seems likely that natural
biofilms are often initiated from pre-formed aggregates.
Despite this, the role of pre-formed aggregates in biofilm
development has, to our knowledge, not yet been ad-
dressed.
In this paper, we have investigated the fate
of pre-aggregated cells during biofilm formation, using
individual-based simulations. Our study shows that an
initial aggregate can have a significant and long-lasting
effect on biofilm spatial structure, even after many gen-
erations of cell growth. Focusing on the role of aggre-
gate shape, we find that, in the absence of competition
for nutrients from surrounding cells, an aggregate that is
initially spread on the surface is favoured over one that is
initially rounded even over long periods of biofilm devel-
opment. This is likely to be because the spread aggregate
has initially a larger surface over which it can absorb nu-
trients, giving it an initial growth advantage that is then
maintained as the biofilm grows.
Strikingly though, our results change qualitatively in
the presence of competition from surrounding, unaggre-
gated cells. When this competition is strong, although
the spread aggregates still grow faster in the early stages
of biofilm development, rounded aggregates become more
successful (produce more progeny) as the biofilm devel-
ops over longer times. This effect appears to arise from a
trade-off between height (as nutrients diffuse from above)
and exposed surface area.
In the absence of competi-
tion, surface area is more important than height, and the
spread aggregate is favoured. However, in the presence
of competition, height becomes more important, since
cells at the top of the aggregate can avoid competing for
nutrients with the surrounding competitors. Since the
rounded aggregate is taller than the spread aggregate,
it gains a "fitness" advantage under conditions of strong
8
competition that is only realised after long times.
Bacterial biofilm formation is a complex phenomenon
which involves a plethora of biological mechanisms in-
cluding cell motility [48], EPS production [49, 50],
metabolic and other phenotypic differentiation [9, 47, 51],
and cell-cell interactions such as quorum sensing [52 -- 54].
In our simulations, almost all of this biological complex-
ity has been neglected; our model takes account only
of nutrient gradients established by cell consumption,
nutrient-dependent growth, and competition among cells
for space. Nevertheless this simplistic approach produces
biologically interesting, and potentially testable, predic-
tions.
In particular our simulations predict that being
initially spread on a surface is a better strategy for a
bacterial aggregate in the absence, but not in the pres-
ence, of competition. Understanding how further bio-
logical complexity might affect this picture would be a
very interesting topic for further work. Another avenue
worth investigating would be the effects of biofilm erosion
and the subsequent detachment of cells. Here, our simu-
lations have not included the effects of fluid flow, which
among other effects, may flatten the biofilm by detaching
protruding cells.
How might aggregates of different shape arise in na-
ture? It is well known that bacterial interactions with
surfaces can vary greatly depending both on the physi-
cal and chemical properties of the surface [55 -- 57], and
on bacterial phenotypes such as EPS production and the
presence of surface appendages. It is therefore very likely
that aggregates formed from bacteria of different taxa or
strains, landing on different surfaces, might adopt dif-
ferent configurations. For example, certain bacteria pro-
duce surfactant which can alter the morphology of a de-
veloping biofilm and allow them to expand over surfaces
more efficiently [58, 59].
Our work has been inspired by the observation that
bacterial aggregates often form in the planktonic phase
[11 -- 13]. Aggregates are known also to form via the
detachment of bacterial clumps from a mother biofilm
[15, 60, 61]; should such aggregates land on a pristine
surface, similar phenomena to those discussed here would
be expected to arise. Moreover, our results could also be
relevant to aggregates that form on the surface itself.
In the classical picture of P. aeruginosa biofilm devel-
opment, individual cells land on a surface, upon which
they migrate and proliferate to form small aggregates
(i.e. microcolonies). Surface-induced motility mecha-
nisms [62, 63] such as twitching [64], crawling [65] and
walking [66] have been implicated in this process. Once
formed, such small aggregates would compete with sur-
rounding cells for nutrients in much the same way as the
pre-formed aggregates that we investigate in this paper.
This study has focused on a single pre-formed aggre-
gate seeding the surface and competing with initially un-
aggregated cells during biofilm formation. To further un-
derstand biofilms in nature, this work should be extended
to investigate competition between multiple aggregates
arranged on the surface, and competition between ag-
9
FIG. 8. Distribution of fittest cells varies with aggregate shape. 2D histograms representing the number of progeny, N , produced
(480 h) by individual bacteria as a function of their initial location in the spread and rounded aggregates in the absence and
presence of competition: (a) θ = 5◦, ρ = 0.0 cell µm−1; (b) θ = 180◦, ρ = 0.0 cell µm−1; (c) θ = 5◦, ρ = 0.5 cell µm−1; (d)
θ = 180◦, ρ = 0.5 cell µm−1. Note that these distributions were averaged over 20 trajectories for each aggregate. Note that the
gradient in the number of progeny is so large that a log scale is used for visualisation purposes.
gregates and mixed strains of bacteria, i.e., strains with
different growth rates.
Recently cooperation within clumps of aggregated cells
has been suggested to be a major stepping stone in the
evolution of multicellularity [26, 27]. Our study thus also
hints that interesting social interactions might arise be-
tween cells within an aggregate. For all aggregate shapes,
we observe heterogeneity in fitness among cells within
the aggregate. This is particularly pronounced for the
rounded aggregate, where cells at the top are strongly
favoured while those in the centre of the aggregate hardly
proliferate. Based on arguments recently put forward
by West and Biernaskie [26, 27], one might predict that
rounded aggregates would be favourable under condi-
tions where cells within the aggregate are closely related,
whereas spread aggregates, in which fitness differences
between cells are less pronounced, might form where cells
are less closely related. This leads to interesting fur-
ther questions, e.g., when a rounded aggregate initiates
biofilm growth, does the majority of cells in the aggre-
gate "sacrifice" their future progeny in favour of their kin
at the top? This idea supports previous suggestions that
height plays a crucial role in competition within biofilms
[32]. While previous work pointed to EPS production
as a means to push progeny cells above the surrounding
competitors [32, 67], our work shows that aggregate for-
mation also provides a means to this end. Such a picture
raises new questions about the evolutionary implications
of bacterial aggregation.
ACKNOWLEDGEMENTS
GM would like to thank Diarmuid Lloyd and Bartek
Waclaw for helpful discussions, and Jan Ulrich Kreft,
Robert Clegg, and Kieran Alden for technical advice
throughout the duration of this study. All authors wish
to thank The Human Frontiers Science Program for fi-
nancial support under Grant Number RGY 0081/2012.
In addition GM and RA would also like to thank the EP-
SRC EP/J007404. Also, KK and TB were supported by
the Lundbeck Foundation. KK was supported by Oticon
Fonden and Knud Højgaards Fond. RJA was supported
by a Royal Society University Research Fellowship.
AUTHOR CONTRIBUTIONS
GM designed and performed the simulation work, and
drafted the manuscript. RA designed and coordinated
the study and helped draft the manuscript. The idea of
investigating the growth of pre-formed aggregates on sur-
faces was conceived by TB, VG, RA, and SG. KK, JH,
YI, and AR all aided in the design of the study and the
biological interpretation of the results. All authors pro-
vided critical evaluation of the study, assisted in editing
the manuscript, and provided approval for submission.
COMPETING INTERESTS
The authors declare no competing interests.
[1] Costerton, J. W., Lewandowski, Z., Caldwell, D. E., Ko-
rber, D. R. & Lappin-Scott, H. M. Microbial biofilms.
Annu. Rev. Microbiol. 49, 711 -- 45 (1995). URL http:
//www.ncbi.nlm.nih.gov/pubmed/8561477.
[2] Costerton, J. W., Montanaro, L. & Arciola, C. R. Biofilm
in implant infections: Its production and regulation. Int.
J. Artif. Organs 28, 1062 -- 1068 (2005).
[3] Fazli, M. et al. Nonrandom distribution of Pseudomonas
aeruginosa and Staphylococcus aureus
in chronic
wounds. J .Clin. Microbiol. 47, 4084 -- 4089 (2009). URL
http://www.pubmedcentral.nih.gov/articlerender.
fcgi?artid=2786634&tool=pmcentrez&rendertype=
abstract.
[4] Bjarnsholt, T. et al. Pseudomonas aeruginosa biofilms in
the respiratory tract of cystic fibrosis patients. Pediatr
Pulmonol 44, 547 -- 558 (2009).
[5] Burmølle, M. et al. Biofilms in chronic infections - a
matter of opportunity - monospecies biofilms in mul-
tispecies infections. FEMS Immunol. Med. Mic. 59,
324 -- 36 (2010). URL http://www.ncbi.nlm.nih.gov/
pubmed/20602635.
[6] Høiby, N., Bjarnsholt, T., Givskov, M., Molin, S. r. &
Ciofu, O. Antibiotic resistance of bacterial biofilms. Int.
J. Antimicrob. Ag. 35, 322 -- 32 (2010). URL http://www.
ncbi.nlm.nih.gov/pubmed/20149602.
[7] Kharazmi, A. Mechanisms involved in the evasion of
the host defence by Pseudomonas aeruginosa. Immunol.
Lett. 30, 201 -- 206 (1991). URL http://www.ncbi.nlm.
nih.gov/pubmed/1757106.
[8] Stewart, P. S. & Franklin, M. J. Physiological heterogene-
ity in biofilms. Nat. Rev. Microbiol. 6, 199 -- 210 (2008).
URL http://www.ncbi.nlm.nih.gov/pubmed/18264116.
[9] Sauer, K., Camper, A. K., Ehrlich, G. D., Costerton,
J. W. & Davies, D. G. Pseudomonas aeruginosa Displays
Multiple Phenotypes during Development as a Biofilm. J.
Bacteriol 184, 1140 -- 1154 (2002).
The develop-
[10] Monds, R. D. & O'Toole, G. A.
mental model of microbial biofilms:
ten years of a
paradigm up for review. Trends Microbiol. 17, 73 -- 87
(2009). URL http://www.sciencedirect.com/science/
article/pii/S0966842X09000055.
[11] Bossier, P. & Verstraete, W. Triggers for microbial ag-
gregation in activated sludge? Appl. Microbiol. Biot. 45,
1 -- 6 (1996). URL http://link.springer.com/10.1007/
s002530050640.
aeruginosa
[12] Alhede, M. et al. Phenotypes of non-attached Pseu-
domonas
surface
attached biofilm. PloS one 6, e27943 (2011). URL
http://www.pubmedcentral.nih.gov/articlerender.
fcgi?artid=3221681&tool=pmcentrez&rendertype=
abstract.
aggregates
resemble
[13] Haaber, J., Cohn, M. T., Frees, D., Andersen, T. r. J.
& Ingmer, H. Planktonic aggregates of Staphylococcus
10
aureus protect against common antibiotics. PloS one
7, e41075 (2012). URL http://www.pubmedcentral.
nih.gov/articlerender.fcgi?artid=3399816&tool=
pmcentrez&rendertype=abstract.
[14] Faruque, S. M. et al. Transmissibility of cholera:
in
vivo-formed biofilms and their relationship to infectivity
and persistence in the environment. Proc. Natl. Acad.
Sci. USA 103, 6350 -- 5 (2006).
URL http://www.
pubmedcentral.nih.gov/articlerender.fcgi?artid=
1458881&tool=pmcentrez&rendertype=abstract.
[15] Hall-Stoodley, L. & Stoodley, P. Biofilm formation
and dispersal and the transmission of human pathogens.
Trends. Microbiol. 13, 7 -- 10 (2005). URL http://www.
ncbi.nlm.nih.gov/pubmed/15639625.
[16] Anton, V., Roug´e, P. & Daff´e, M. Identification of the
sugars involved in mycobacterial cell aggregation. FEMS
Microbiology Letters 144, 167 -- 170 (1996).
[17] Monier, J. & Lindow, S. E. Frequency , Size , and Lo-
calization of Bacterial Aggregates on Bean Leaf Surfaces.
Appl. Environ. Microb. 70, 346 -- 355 (2004).
[18] Lens, P., Moran, A. P., Mahony, T., Stoodley, P. &
O'Flaherty, V. Biofilms in medicine, industry and en-
vironmental biotechnology: characteristics, analysis and
control, vol. 1 (IWA Publishing, London, 2003).
[19] Kreft, J.-U., Picioreanu, C., Wimpenny, J. W. T. &
van Loosdrecht, M. C. M.
Individual-based modelling
of biofilms. Microbiology 147, 2897 -- 912 (2001). URL
http://www.ncbi.nlm.nih.gov/pubmed/11700341.
[20] Xavier, J. B., Picioreanu, C. & van Loosdrecht, M. C. M.
A framework for multidimensional modelling of activity
and structure of multispecies biofilms. Environ. Micro-
biol. 7, 1085 -- 1103 (2005). URL http://www.ncbi.nlm.
nih.gov/pubmed/16011747.
[21] Alpkvist, E., Picioreanu, C., Loosdrecht, M. C. M. V.
& Heyden, A. Three-Dimensional Biofilm Model With
Individual Cells and Continuum EPS Matrix. Biotechnol.
Bioeng. 94, 961 -- 979 (2006).
[22] Kreft, J.-U. Biofilms promote altruism. Microbiology
150, 2751 -- 2760 (2004). URL http://www.ncbi.nlm.
nih.gov/pubmed/15289571.
[23] Nadell, C. D., Foster, K. R. & Xavier, J. a. B. Emergence
of spatial structure in cell groups and the evolution of
cooperation. Plos. Comp. Biol. 6, e1000716 (2010). URL
http://www.pubmedcentral.nih.gov/articlerender.
fcgi?artid=2841614&tool=pmcentrez&rendertype=
abstract.
[24] Tuson, H. H. & Weibel, D. B.
Soft matter 9,
Bacteria-surface
4368 -- 4380 (2013).
http://www.sciencemag.org/content/256/
interactions.
URL
5056/495.shorthttp://www.ncbi.nlm.nih.gov/
pubmed/23930134http://www.pubmedcentral.nih.
gov/articlerender.fcgi?artid=3733390&tool=
pmcentrez&rendertype=abstract.
[25] Israelachvili,
face Forces
http://www.sciencedirect.com/science/article/
pii/B9780123751829100119.
J. N.
(Elsevier,
Intermolecular
third edn.
2011),
and Sur-
URL
[26] Biernaskie, J. M. & West, S. A. Cooperation, clumping
and the evolution of multicellularity. Proc. R. Soc. B
(2015).
[27] West,
S. A.,
in individuality.
Fisher, R. M., Gardner, A.
transi-
Proc. Natl. Acad. Sci.
(2015).
URL http:
& Kiers, E. T.
tions
USA 112,
//www.pnas.org/content/112/33/10112.abstract.
http://www.pnas.org/content/112/33/10112.full.pdf.
10112 -- 10119
evolutionary
Major
[28] Hermanowicz, S. W. A Simple 2D Biofilm Model Yields
a Variety of Morphological Features. Math. Biosci. 169,
1 -- 14 (2001).
[29] Stewart, P. S. Diffusion in Biofilms. J. Bacteriol 185,
1485 -- 1491 (2003).
[30] Dockery, J. & Klapper,
Finger Formation in
Biofilm Layers.
SIAM J. Appl. Math. 62, 853 --
869 (2001). URL http://epubs.siam.org/doi/abs/10.
1137/S0036139900371709.
I.
[31] Douezan, S. et al. Spreading Dynamics and Wetting
Transition of Cellular Aggregates. Proc. Natl. Acad.
Sci. USA 108, 7315 -- 7320 (2011). URL http://www.
pubmedcentral.nih.gov/articlerender.fcgi?artid=
3088609&tool=pmcentrez&rendertype=abstract.
[32] Xavier, J. B. & Foster, K. R. Cooperation and conflict
in microbial biofilms. Proc. Natl. Acad. Sci. USA 104,
876 -- 881 (2007).
URL http://www.pubmedcentral.
nih.gov/articlerender.fcgi?artid=1783407&tool=
pmcentrez&rendertype=abstract.
[33] Lardon, L. A. et al.
iDynoMiCS: Next-Generation
Individual-Based Modelling of Biofilms. Environ. Micro-
biol. 13, 2416 -- 34 (2011). URL http://www.ncbi.nlm.
nih.gov/pubmed/21410622.
[34] Xavier, J. B., Picioreanu, C., Rani, S. A., van Loosdrecht,
M. C. M. & Stewart, P. S. Biofilm-control strategies
based on enzymic disruption of the extracellular poly-
meric substance matrix -- a modelling study. Microbiol-
ogy (Reading, England) 151, 3817 -- 3832 (2005). URL
http://www.ncbi.nlm.nih.gov/pubmed/16339929.
[35] Characklis, W. G. & Marshall, K. C. Biofilms (Wiley
Interscience, New York, 1990).
11
[41] Shedlovsky, T. Electrochemistry in Biology and Medicine
(John Wiley and Sons, New York, 1955).
[42] Picioreanu, C., Van Loosdrecht, M. C. M. & Heijnen,
J. J. Mathematical modeling of biofilm structure with a
hybrid differential- discrete cellular automaton approach.
Biotechnol. Bioeng. 58, 101 -- 116 (1998).
[43] Van Loosdrecht, M. C. M., Heijnen, J. J., Eberl, H.,
Kreft, J. & Picioreanu, C. Mathematical modelling of
biofilm structures. Anton. Leeuw. Int. J. G. 81, 245 -- 256
(2002).
[44] Wentland, E. J., Stewart, P. S., Huang, C.-T. &
McFeters, G. A. Spatial variations in growth rate within
klebsiellapneumoniae colonies and biofilm. Biotechnol.
Prog. 12, 316 -- 321 (1996). URL http://dx.doi.org/10.
1021/bp9600243.
[45] .
[46] Monier, J. M. & Lindow, S. E. Differential survival
of solitary and aggregated bacterial cells promotes
aggregate formation on leaf surfaces.
Proc. Natl.
Acad. Sci. USA 100, 15977 -- 15982 (2003). URL http:
//www.pubmedcentral.nih.gov/articlerender.fcgi?
artid=307678&tool=pmcentrez&rendertype=abstract.
[47] Stoodley, P., Sauer, K., Davies, D. G. & Costerton, J. W.
Biofilms as complex differentiated communities. Annu.
Rev. Microbiol. 56, 187 -- 209 (2002). URL http://www.
ncbi.nlm.nih.gov/pubmed/12142477.
[48] Klausen, M. et al. Biofilm formation by Pseudomonas
aeruginosa wild type, flagella and type IV pili mutants.
Mol. Microbiol. 48, 1511 -- 1524 (2003). URL http://www.
ncbi.nlm.nih.gov/pubmed/12791135.
[49] Flemming, H.-C., Neu, T. R. & Wozniak, D. J. The EPS
matrix: the "house of biofilm cells". J. Bacteriol 189,
7945 -- 7947 (2007). URL http://www.pubmedcentral.
nih.gov/articlerender.fcgi?artid=2168682&tool=
pmcentrez&rendertype=abstract.
[50] Flemming, H.-C. & Wingender, J. The biofilm matrix.
Nat. Rev. Microbiol. 8, 623 -- 633 (2010).
[51] Watnick, P. & Kolter, R. Biofilm , City of Microbes. J.
Bacteriol 182, 2675 -- 2679 (2000).
[52] Davies, D. G. The Involvement of Cell-to-Cell Signals
Science
in the Development of a Bacterial Biofilm.
280, 295 -- 298 (1998). URL http://www.sciencemag.
org/cgi/doi/10.1126/science.280.5361.295.
[53] Miller, M. B. & Bassler, B. L. Quorum sensing in bacte-
[36] Bailey, J. E. & Ollis, D. F. Biochemical Engineering
ria. Annu. Rev. Microbiol. 55, 165 -- 199 (2001).
Fundamentals (McGraw-Hill, New York, 1986).
[37] Bakke, R., Trulear, M. G., Robinson, J. A. & Characklis,
W. G. Activity of Pseudomonas aeruginosa in Biofilms:
Steady State. Biotechnol. Bioeng. 26, 1418 -- 1424 (1984).
[38] Robinson, J. A., Trulear, M. G. & Characklis, W. G.
Cellular Reproduction and Extracellular Polymer Forma-
tion by Pseudomonas aeruginosa in Continuous Culture.
Biotechnol. Bioeng. 26, 1409 -- 1417 (1984).
[39] Beyenal, H., Chen, S. N. & Lewandowski, Z. The dou-
ble substrate growth kinetics of Pseudomonas aeruginosa.
Enzyme. Microb. Tech. 32, 92 -- 98 (2003).
in biofilms with new numerical
[40] Horn, H., Neu, T. R. & Wulkow, M. Modelling
the structure and function of extracellular polymeric
substances
tech-
niques. Water Sci. Technol. 43, 121 -- 127 (2001). URL
http://www.ncbi.nlm.nih.gov/pubmed/11381957$\
delimiter"026E30F$nhttp://www.ncbi.nlm.nih.gov/
entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=
Citation&list_uids=11381957.
[54] Diggle, S. P., Griffin, A. S., Campbell, G. S. & West,
S. A. Cooperation and conflict in quorum-sensing bacte-
rial populations. Nature 450, 411 -- 414 (2007).
[55] Dorobantu, L. S., Bhattacharjee, S., Foght, J. M. & Gray,
M. R. Analysis of force interactions between AFM tips
and hydrophobic bacteria using DLVO theory. Langmuir
25, 6968 -- 6976 (2009).
[56] Harimawan, A., Rajasekar, A. & Ting, Y. P. Bacteria
attachment to surfaces - AFM force spectroscopy and
physicochemical analyses. Journal of Colloid and Inter-
face Science 364, 213 -- 218 (2011).
[57] Wright, C. J., Shah, M. K., Powell, L. C. & Armstrong,
I. Application of AFM from Microbial Cell to Biofilm.
Scanning 32, 134 -- 149 (2010).
[58] Davey, M. E., Caiazza, N. C. & O'Toole, G. A. Rham-
nolipid surfactant production affects biofilm architecture
in Pseudomonas aeruginosa PAO1. J. bacteriol. 185,
1027 -- 36 (2003).
URL http://www.pubmedcentral.
nih.gov/articlerender.fcgi?artid=142794{&}tool=
pmcentrez{&}rendertype=abstract.
[63] Henrichsen, J.
Bacterial surface translocation:
12
[59] Angelini, T. E., Roper, M., Kolter, R., Weitz, D. A.
& Brenner, M. P. Bacillus subtilis spreads by surf-
Proc. Natl. Acad.
ing on waves of
Sci. USA 106, 18109 -- 18113 (2009).
URL http:
//www.pnas.org/content/106/43/18109.abstract.
http://www.pnas.org/content/106/43/18109.full.pdf.
surfactant.
[60] Wilson, S., Hamilton, M. a., Hamilton, G. C., Schumann,
M. R. & Stoodley, P. Statistical quantification of de-
tachment rates and size distributions of cell clumps from
wild-type (PAO1) and cell signaling mutant (JP1) Pseu-
domonas aeruginosa biofilms. Appl. Environ. Microb. 70,
5847 -- 5852 (2004).
[61] Stoodley, P., Hall-Stoodley, L. & Lappin-Scott, H. M.
Detachment, surface migration, and other dynamic be-
havior in bacterial biofilms revealed by digital time-lapse
imaging. Method Enzymol. 337, 306 -- 19 (2001). URL
http://www.ncbi.nlm.nih.gov/pubmed/11398439.
[62] Harshey, R. M. Bacterial Motility on a Surface: Many
Ways to a Common Goal. Annu. Rev. Microbiol. 57,
249 -- 73 (2003). URL http://www.ncbi.nlm.nih.gov/
pubmed/14527279.
a
Bacteriol. Rev. 36,
URL http://www.pubmedcentral.
survey and a classification.
478 -- 503 (1972).
nih.gov/articlerender.fcgi?artid=408329&tool=
pmcentrez&rendertype=abstract.
[64] Burrows, L. L. Pseudomonas aeruginosa Twitching
Motility: Type IV Pili in Action. Annu. Rev. Microbiol.
66, 493 -- 520 (2012).
[65] Conrad, J. C. et al.
Flagella and Pili-Mediated
Near-Surface Single-Cell Motility Mechanisms in P.
aeruginosa. Biophys. J. 100, 1608 -- 1616 (2011). URL
http://www.sciencedirect.com/science/article/
pii/S0006349511002372.
[66] Gibiansky, M. L. et al. Bacteria use type IV pili to walk
upright and detach from surfaces. Science (New York,
N.Y.) 330, 197 (2010).
[67] Kim, W., Racimo, F., Schluter, J., Levy, S. B. & Foster,
K. R. Importance of positioning for microbial evolution.
Proc. Natl. Acad. Sci. USA 111, E1639 -- E1647 (2014).
URL http://www.ncbi.nlm.nih.gov/pubmed/24715732.
Supplementary Material
Shaping the Growth Behaviour of Biofilms
Initiated from Bacterial Aggregates
S1 Generating 2D bacterial aggregates
To generate bacterial aggregates of the appropriate shape, circular segments
were extracted from simulation configurations of pre-grown biofilms as illus-
trated in Figures S1 and S2. Figure S1(a) shows a circle (red) with radius R
and segmental region defined by the white area between the black horizontal
line and the perimeter of the circle. The "surface-aggregate angle", θ, defines
the angle between the black horizontal line (surface) and the tangent line of
the circle at the point of contact with the surface.
The area, A, of a circular segment of radius R and with surface-aggregate
angle of θ, is given by [1]
(S1)
(2θ − sin2θ)R2.
1 2
A =
The corresponding arc length, s, of an aggregate defines the initial interface
between the bacterial aggregate and the surrounding nutrient medium, and
is computed according to
(S2)
To ensure our aggregates contained the same number of cells (∼100), the
area in Equation S1 was fixed at 1257 µm2.
The procedure for generating the starting aggregate configurations was
s = 2Rθ.
performed as follows:
1
arXiv:1506.08168v2 [physics.bio-ph] 24 Nov 2015
1. An in silico biofilm was grown for a simulation time of 7 days. This
simulation was initiated with 200 cells randomly distributed on the
surface and a nutrient concentration of 5.4×10−2 g L−1 was used. A
typical biofilm resulting from such a simulation was approximately 800
µm in height with some finger-like projections such as those in Figure
S2.
2. A point, P , was defined in the middle of the biofilm (Figure S1 and
S2). The point P is the reference point from which the origin, O, of
a circle is computed. The point P was selected such that the density
of cells within any segmental area was relatively uniform. Bacterial
aggregates of shapes θ = 5, 10, 15, 20, 25, ....., 180o with fixed area of
1257 µm2 were then generated as in Figure S2.
3. For each θ, the radius of the corresponding circle was computed by
solving Equation S1 for R. Using the radius, R, of this new circle, we
computed the distance, X, to move the origin of this new circle above
or below the point P according to
where
X = R sinα,
α = 90 − θ.
(S3)
(S4)
4. With the new origin and radius defined, we then extracted all bacteria
within the radial distance from O. Note those bacteria located below
the horizontal line (blue region) were excluded. The corresponding
segment gave rise to a bacterial aggregate with a shape determined by
the value of θ.
5. The vertical component of the point P was then subtracted from all
bacterial coordinates so that the base of the aggregate was located on
the surface as shown in Figure 1(a) of the main manuscript.
S2 Generating configurations of competing "sin-
gle cells" in 2D
The width of the surface available to the surrounding competitor cells de-
pends on the surface coverage of the aggregate. As surface coverage of the
aggregate increases, θ → 0, the amount of available surface either side of the
aggregate decreases. Surrounding cells were inserted at random positions
2
Supplementary Figure S1: Schematic representation of the geometry of ag-
gregates in (a) 2D and (b) 3D. The light blue region represents the area not
included when computing cells within a distance R from the origin, O. The
angle θ is the contact angle of the aggregate with the surface.
Supplementary Figure S2: Generating aggregates with different geometries
but containing the same number of cells. Using pre-grown 2D biofilm con-
figurations, green, aggregates of the desired shape can be generated using
the geometric relationships from Figure S1(a). (a) Schematic showing how a
spread aggregate is generated. (b) Schematic showing how a more rounded
aggregate is generated.
3
on the surface excluding the region occupied by the aggregate. To ensure a
constant density of cells surrounding the aggregate, the number of cells in
these regions were varied according to the width of surface available.
S3 Generating 3D aggregates
To generate aggregates in our 3D simulations, we performed an analogous
process to that used for the 2D case. Configurations of spherical caps were
transplanted from pre-grown biofilms using the geometrical relationships il-
lustrated in Figure S1(b). The volume of a spherical cap is [2]
(S5)
(S6)
(S7)
πR3(2 − 3sinα + sin3α),
1 3
Vcap =
and the corresponding surface area of the cap is given by
where the height, h, is given by
S = 2πRh,
h = R − X
A spherical cap volume, Vcap, of 5575 µm3 was used for all values of θ, giving
approximately 118 bacteria in each aggregate.
S4 Generating configurations of competing "sin-
gle cells" in 3D
To surround the 3D aggregates with neighbours, we seeded the surface sur-
rounding the aggregate with single cells. The number of these competing
cells per square µm gave the density of cells on the surface. Cells that lay in
the region defined by the radius of the spherical cap, a, were removed from
this configuration. The aggregate configuration and the surrounding cell
configuration were then combined to produce the initial state of the system.
S5 Multiple simulation runs in 2D and 3D
For our 2D simulations, four different aggregate configurations were gener-
ated for each value of θ, and for each of these configurations, 5 simulations of
480 hours were performed using a different initial distribution of surrounding
cells. This gave rise to a total of 20 simulations for each value of θ.
4
In 3D, only the spread aggregate, θ = 5o, and the rounded aggregate, θ =
180o, were simulated. Three different initial configurations were generated
for each value of θ, and for each of these configurations 3 simulations of 72
hours were performed, each using different initial distributions of surrounding
single cells. This gave rise to a total of 9 simulations for each value of θ.
Note that, in both the 2D and 3D simulations, the competitor cells on
the surface and the aggregated cells are identical in their growth parameters
(Table 1, main manuscript).
S6 Exploration of Biofilm Structure at Low
and High Nutrient Concentration With-
out Aggregates
To investigate the effect of nutrient concentration in general in our simu-
lations, we simulated 2D biofilms in the absence of initial aggregates, at
several different nutrient concentrations and various densities of cells, which
were initially located at random across the surface. Figures S3 (a) and (b)
show the resulting biofilms formed after 120 h at a nutrient concentration of
5.4 × 10−2g L−1 for seeder cell densities of 0.01 cell µm−1 and 0.5 cell µm−1.
Clearly, the biofilms formed at this nutrient concentration are significantly
thicker than those formed at lower nutrient concentration (Figures S3(c) to
(f)). Comparing (a) and (b), we see that a low density of seeder cells leads
to a less uniform biofilm morphology with finger-like projections.
After 120 h, at low nutrient concentrations (5.4×10−3g L−1), the biofilms
remain very thin (Figures S3(c) and (d)). At this low concentration, the
initial number of seeder cells has a marked effect on the final structure of
the biofilm. A high density of seeder cells yields a biofilm with a uniform
morphology whereas a low density of seeder cells yields a biofilm with a
spatial structure that is clearly dependent on the initial configuration of the
seeder cells. Figures S3(e) and (f) show that this effect persists at long times.
From Figures S3(a) and (e), it is clear that both longer times and/or
higher nutrient concentration lead to larger biofilms. Selecting a bulk nutrient
concentration and timescale to investigate the shape dependent fitness of
aggregated clusters of bacteria is therefore non-trivial. In our simulations,
we used the lower nutrient concentration of 5.4 × 10−3g L−1 (see Table 1 in
the main manuscript), comparable with previous biofilm simulations [3, 4, 5].
We also chose to run the simulations for 480 h in order to explore the long
term growth dynamics, and to generate populations that were large enough
to obtain good statistics. Timescales of 480 h have also been used in previous
5
Supplementary Figure S3: Nutrient concentration and initial density of
seeder cells affect biofilm morphology in the absence of initial aggregates:
(a) Nutrient concentration = 5.4 × 10−2g L−1, Density of seeder cells = 0.01
cell µm−1. (b) Nutrient concentration = 5.4 × 10−2g L−1, Density of seeder
cells = 0.5 cell µm−1. (c) Nutrient concentration = 5.4× 10−3g L−1, Density
of seeder cells = 0.01 cell µm−1. (d) Nutrient concentration = 5.4 × 10−3g
L−1, Density of seeder cells = 0.5 cell µm−1.
(e) Nutrient concentration
= 5.4× 10−3g L−1, Density of seeder cells = 0.01 cell µm−1, 480 h. (f) Nutri-
ent concentration = 5.4× 10−3g L−1, Density of seeder cells = 0.5 cell µm−1,
480h.
simulation studies [3, 4]. After 480 h growth using a bulk concentration of
5.4 × 10−3g L−1, our simulations produce biofilms that are approximately
200-300 µm in height (Figures S3(e) and (f)).
6
Supplementary Figure S4: Without competing cells on the surface, the func-
tional form of the N/N0 curves remains relatively constant over time. Note,
however that the scale does change because more cells are produced as time
increases: (a) 56 h. (b) 120 h. (c) 240 h. (d) 480 h. Error bars represent the
standard deviation from 20 simulations.
S7 Exploration of Aggregate Fate at Differ-
ent Times
In the main manuscript, we computed N/N0, as a function of the aggregate
surface angle, θ, after 480 h. Here, we show that qualitatively similar results
are obtained regardless of the time at which fitness is measured. Figures S4
shows N/N0 as function of θ in the absence of competing cells (no red cells
in Figure 1(a) of the main manuscript) at 4 different times. Evidently, the
functional form of the curves does not change significantly with time.
In the highly competitive regime where we include competitor cells (red
cells in Figure 1 of main manuscript) at a density of 0.5 cells µm−1, the
situation is more complex. Here, the functional form of the curves changes
during the course of the simulation (Figure S5). For early times, the curves
are similar to the case without competition, suggesting that shape-dependent
nutrient access is the still dominant factor in the early stages. As time
7
(b) 120 h
(d) 480 h
22
20
18
16
0
N / N
14
12
145
140
135
130
125
120
115
110
105
0
N / N
(a) 56 h
(c) 240 h
10
9
8
7
6
5
55
50
45
40
35
0
N / N
0
N / N
0 20 40 60 80 100 120 140 160 180
0 20 40 60 80 100 120 140 160 180
θ [
o
]
θ [
o
]
Supplementary Figure S5: With competing cells on the surface at a density
of 0.5 cell µm−1, the functional form of the N/N0 curve changes over time:
(a) 56 h. (b) 120 h. (c) 240 h. (d) 480 h. Error bars represent the standard
deviation from 20 simulations.
increases, the rounded aggregates become increasingly "fitter" relative to
the spread aggregates, until at 480 h, all aggregates with a surface aggregate
angle greater than 75o are fitter than the spread aggregate. From this we
can conclude that the initial height advantage of the rounded aggregate only
becomes important at longer times.
S8 Difference in Fate Between Spread and
Rounded Aggregates at Higher Nutrient
Concentration
Here, we investigate the fate of the spread and round aggregate at a higher
nutrient concentration, 1 × 10−2 gL−1. Figure S6 shows that our conclusions
remain valid at this nutrient concentration: at low density of seeder cells there
is no difference in fate between the spread and rounded aggregates, whereas
at higher density of initial neighbours, the rounded aggregate is favoured
8
(b) 120 h
(d) 480 h
12
11
10
9
8
7
6
50
48
46
44
42
40
38
36
34
32
30
28
0
N / N
0
N / N
(a) 56 h
(c) 240 h
6.5
6
5.5
5
4.5
4
3.5
3
0
N / N
22
21
20
19
18
17
16
15
14
13
0
N / N
0 20 40 60 80 100 120 140 160 180
0 20 40 60 80 100 120 140 160 180
θ [
o
]
θ [
o
]
Supplementary Figure S6: At higher nutrient concentration, 1 × 10−2gL−1,
rounded aggregates are more favourable for growth. At low density of seeder
cells (0.01 cell µm−1) there is no significant difference in N/N0 between the
spread and rounded aggregates. At high density (0.5 cell µm−1), the rounded
aggregate is again produces more progeny per initial cells relative to that of
the spread. Note that these data were generated from simulations represent-
ing 480 h of biofilm growth.
over its initially more spread counterpart. Note that these simulations were
again run for 480 h.
S9 The Effect of Aggregate Shape on the Growth
of the Active Layer
In the main manuscript we discussed cell growth rate heterogeneity within
the biofilm and the presence of a well-defined active layer in the growing
aggregate. The depth of the active later has been shown to be very important
in determining biofilm structure and the degree of segregation amongst cell
groups within the community [6].
It has also been discussed in extensive
mathematical detail by Dockery and Klapper [7]. Here we look at the growth
9
P = 0,5342
***
P < 0,0001
Spread
Rounded
0.0 1
0.5
Density of Competitor Cells [cell um-1]
200
150
100
50
0
N/N0
Supplementary Figure S7: Dynamics of the growing layer determine aggre-
gate growth behaviour. The number of growing cells in the spread and
rounded aggregates at various competitive regimes. For clarity the error
bars, representing the standard deviations, are only shown for the final data
points. The standard deviations at these points are maximal.
of the active layer.
In the analysis that follows, we simply define the active layer as the
region of the growing aggregate in which all cells grow with a growth rate
greater than 1 fg h−1 (see Figure 5 of main manuscript), and we assess the
number of cells in this layer. This choice is arbitrary; what is important is
the qualitative difference in behaviour between the aggregate shapes.
Figure S7(a) shows the number of active cells within the spread (θ =
5o) and rounded (θ = 180o) aggregates at varying degrees of competition;
dictated by the density of surrounding unaggregated cells on the surface. In
the absence of competition, the number of active cells within the spread and
rounded aggregates increases with time. Initially the number of active cells
in the spread aggregate grows faster than that for its rounder counterpart,
however for t > 100 h, the number of active cells in both aggregates increases
with similar rates (slope of red curves). This linear increase determines the
super-linear increase in the corresponding aggregate populations in Figure
1(b) of the main manuscript: a linear increase in the number of active cells
translates to a super-linear increase in the total aggregate population.
In all competitive regimes, the number of cells in the growing layer of the
spread aggregates becomes constant after ∼100 h. A constant population of
10
0
0
50 100 150 200 250 300 350 400 450 500
time [h]
Spread, no competition
Rounded, no competition
Spread, low competition
Rounded, low competition
Spread, high competition
Rounded, high competition
5000
4500
4000
3500
3000
2500
2000
1500
1000
500
number of cells in active layer, N
active cells in the growing layer translates to linear growth behaviour of the
total aggregate population (compare solid blue curve with the dashed red
curve in Figure 1(b) of the main manuscript). For the rounded aggregates,
this transition to the constant regime does not occur within the timescales
of our simulations, and the layer continues to grow monotonically.
Comparing the spread and rounded aggregates (solid curves with dashed
curves), with competition, we see that growth of the active layer always in-
creases faster for the spread aggregate before levelling off after 100 h. How-
ever, later in the simulation, the number of active cells in the rounded ag-
gregates becomes greater. The time at which this crossover occurs increases
with increasing levels of competition. The increasing growth of the active
layer in the rounded aggregate at high competition, ρ = 0.5 cell µm−1, ex-
plains why its number of progeny becomes larger than that of the spread
aggregate in Figure 1(b), and why its structures tend to fan outwards at the
top (Figure 4(d)).
S10 Aggregate Simulations in 3D
The results presented in the main manuscript are for 2D simulations, for
reasons of computational feasibility. To validate our results further, we also
ran some simulations in 3D. Figure S8 shows simulation snapshots after 72
h for the spread and rounded aggregates in the presence of low and high
densities of competitor cells. As in our 2D simulations (Figures 3(b) and
(d) of the main manuscript), the rounded aggregate persists as a dominant
structural feature at both low and high densities of competitor cells. The
spread aggregate on the other hand becomes swamped by the surrounding
competitors and after 72 h is structurally indistinguishable from the rest
of the biofilm (as in our 2D simulations, Figures 3(b) and (d) of the main
manuscript).
N/N0 for the spread and rounded aggregates are plotted in Figure S9 and
is analogous to the data shown in Figures 8 of the main manuscript for the
2D simulations. As in the 2D case, the number of progeny per initial cell
of the aggregates decreases with increasing density of surrounding cells. In
the absence of competition, the spread aggregate again produces significantly
more progeny than the rounded aggregate, however as the competition on the
surface is increased, the number of progeny produced by cells in the rounded
aggregate increases relative to that of the spread.
11
Supplementary Figure S8: 3D simulation snapshots showing initial and final
configurations of the spread and rounded aggregates at low and high density
of surrounding cells. (a) Spread aggregate, Density of seeder cells = 0.002
cell µm−2. (b) Spread aggregate, Density of seeder cells = 0.03 cell µm−2. (c)
Rounded aggregate, Density of seeder cells = 0.002 cell µm−2. (d) Rounded
aggregate, Density of seeder cells = 0.03 cell µm−2.
12
Supplementary Figure S9: N/N0 of the rounded aggregate with respect to
the spread aggregate increases with the density of neighbouring cells on the
surface.
In 3D, the trend is similar to that observed in 2D. The spread
aggregate is significantly favoured when there is no competition on the surface
however with increasing density the rounded aggregate is more favourable for
growth.
13
**
P = 0,0081
Spread
Rounded
P = 0,122
***
P < 0,0001
***
P < 0,0001
0
0.0 0 2
0.0 0 4
0.0 3
Density of Competitor Cells [cell um-1]
500
400
300
200
100
0
N/N0
References
[1] Weisstein,
MathWorld -- A
http://mathworld.wolfram.com/CircularSegment.html
Resource.
Web
W.
E.
Wolfram
"Circular
Segment".
Available
From
at:
[2] Weisstein E. W. "Spherical Cap". From MathWorld -- A Wolfram Web Re-
source. Available at: http://mathworld.wolfram.com/SphericalCap.html
[3] Xavier, J. B., Picioreanu, C., Rani, S. A., van Loosdrecht, M. C. M. ,
Stewart, P. S., Biofilm-control strategies based on enzymic disruption of
the extracellular polymeric substance matrix -- a modelling study, Microbiol.
151, 3817-3832 (2005).
[4] Xavier, J. B., Foster, K. R., Cooperation and conflict in microbial
biofilms. Proc. Natl. Acad. Sci. USA 104, 876-881 (2007).
[5] Lardon, L. A., Merkey, B. V., Martins, S., Dtsch, A., Picioreanu. C.,
iDynoMiCS: Next-Generation Individual-
Kreft, J.-U., & Smets, B.F.,
Based Modelling of Biofilms. Environ. Microbiol. 13, 2416-2434 (2011)
[6] Nadell, C. D., Foster, K. R., & Xavier, J.B., Emergence of spatial struc-
ture in cell groups and the evolution of cooperation. PLoS Comput. Biol.
6, e1000716 (2010)
[7] Dockery, J., & Klapper, I., Finger formation in biofilm layers. SIAM J.
Appl. Math 62, 853-869 (2001).
14
|
1910.03941 | 1 | 1910 | 2019-10-07T14:01:58 | High Sensitivity and Specificity Biomechanical Imaging by Stimulated Brillouin Scattering Microscopy | [
"physics.bio-ph",
"physics.optics"
] | Noncontact label-free biomechanical imaging is a crucial tool for unraveling the mechanical properties of biological systems, which play critical roles in the fields of engineering, physics, biology and medicine; yet, it represents a significant challenge in microscopy. Spontaneous Brillouin microscopy meets this challenge, but often requires long acquisition times or lacks high specificity for detecting biomechanical constituents with highly overlapping Brillouin bands. We developed stimulated Brillouin scattering (SBS) microscopy that provides intrinsic noncontact biomechanical contrast and generates mechanical cross-sectional images inside large specimens, with high mechanical specificity and pixel dwell times that are >10-fold improved over those of spontaneous Brillouin microscopy. We used SBS microscopy in different biological applications, including the quantification of the high-frequency complex longitudinal modulus of the pharyngeal region of live wild-type Caenorhabditis elegans nematodes, imaging of the variations in the high-frequency viscoelastic response to osmotic stress in the head of living worms, and in vivo mechanical contrast mesoscopy of developing nematodes | physics.bio-ph | physics | High Sensitivity and Specificity Biomechanical Imaging by
Stimulated Brillouin Scattering Microscopy
Itay Remer1,2, Netta Shemsesh3, Anat Ben-Zvi3 & Alberto Bilenca1,4*
1Biomedical Engineering Department, Ben-Gurion University of the Negev, 1 Ben Gurion Blvd, Be'er-Sheva 84105,
Israel
2Agilent Research Laboratories, 94 Shlomo Shmeltzer Road, Petach Tikva 4970602, Israel
3Department of Life Sciences, Ben-Gurion University of the Negev, 1 Ben Gurion Blvd, Be'er Sheva 84105, Israel
4Ilse Katz Institute for Nanoscale Science and Technology, Ben-Gurion University of the Negev,
1 Ben Gurion Blvd, Be'er-Sheva 84105, Israel
e-mail: *[email protected]
Abstract
Noncontact label-free biomechanical imaging is a crucial tool for unraveling the mechanical
properties of biological systems, which play critical roles in the fields of engineering, physics,
biology and medicine; yet, it represents a significant challenge in microscopy. Spontaneous
Brillouin microscopy meets this challenge, but often requires long acquisition times or lacks high
specificity for detecting biomechanical constituents with highly overlapping Brillouin bands. We
developed stimulated Brillouin scattering (SBS) microscopy that provides intrinsic noncontact
biomechanical contrast and generates mechanical cross-sectional images inside large specimens,
with high mechanical specificity and pixel dwell times that are >10-fold improved over those of
spontaneous Brillouin microscopy. We used SBS microscopy in different biological applications,
including the quantification of the high-frequency complex longitudinal modulus of the pharyngeal
region of live wild-type Caenorhabditis elegans nematodes, imaging of the variations in the high-
frequency viscoelastic response to osmotic stress in the head of living worms, and in vivo
mechanical contrast mesoscopy of developing nematodes.
Main Text
Label-free biomechanical imaging has long used a variety of techniques, including atomic-force
microscopy, multiphoton microscopy, and optical coherence elastography1 -- 6, that obtain
mechanical images with high spatial resolution, but require contact or external mechanical
stimulation of the sample. Spontaneous Brillouin microscopy7 -- 20 circumvents these requirements
by measuring the so-called Brillouin shifts B and linewidths , which are the frequency shifts
and linewidths of light backscattered inelastically from gigahertz-frequency longitudinal acoustic
phonons characteristic to the different viscoelastic constituents of the material. However,
spontaneous Brillouin microscopy often demands long acquisition times due to the low efficiency
of spontaneous Brillouin scattering in biological matter, or suffers from limited mechanical
specificity because of the relatively low spectrometer resolution of spontaneous Brillouin
microscopes, making it difficult to specifically detect biomechanical constituents with highly
overlapping Brillouin bands.
1
Here we introduce stimulated Brillouin scattering (SBS) as a process to significantly enhance the
acquisition rate and mechanical specificity of Brillouin microscopy. SBS is a photon-phonon
scattering process analogous to stimulated Raman scattering but which involves lower-frequency
acoustic phonons. SBS was first observed in 196421 and has widely been used in spectroscopic
research22 -- 29. While in spontaneous Brillouin scattering a single laser beam at a frequency 1
illuminates the sample to produce the entire Brillouin spectrum around the Stokes and anti-Stokes
frequencies ω1 B, in SBS counter narrowband laser pump and probe beams at 1 and 2 overlap
inside the sample to efficiently interact with a particular longitudinal acoustic phonon of
vibrational frequency B (Fig. 1a). When the frequency of the probe beam at 2 is scanned around
S experiences a stimulated Brillouin gain (SBG) via
the Stokes frequency ω1 -- B, its intensity I2
wave resonance, whereas I1 the intensity of the pump beam at 1 shows a stimulated Brillouin loss
(SBL), as depicted in Fig. 1b. The opposite occurs when 2 is scanned around the anti-Stokes
frequency ω1 + B. No gain or loss arise at 2 otherwise. Consequently, unlike spontaneous
Brillouin scattering, SBG or SBL enables Brillouin spectrum measurements that are free of elastic
background interference and that do not trade spectrometer resolution for acquisition time, or vice
versa25 -- 28.
The SBG or SBL spectrum is given by G() = g() l I1, where is the crossing efficiency
of the pump and probe beams in the sample, g() is the SBS gain or loss factor well represented
by a lorentzian line shape, l is the interaction length of the two counter laser beams in the sampled
23. As for the spontaneous Brillouin scattering
volume, and I1 is the intensity of the pump beam at 1
spectrum7, 10, the close relationship between the SBG or SBL spectrum and the complex
longitudinal modulus of the probed volume allows to locally extract the high-frequency
viscoelastic response of the medium. This relationship is described by M* = (1/2n)2 B
2
(1 + iB), where M* is the complex longitudinal modulus, 1 is the wavelength of the beam
at 1, and n and are the refractive index and the mass density of the medium, respectively7, 30.
Similar to other nonlinear optical techniques31, SBS inherently provides optical sectioning in three
dimensions owing to the nonlinearity of SBS in the total irradiance intensity.
SBS has recently been demonstrated for Brillouin microscopy of liquids using tunable single-
frequency lasers in a frequency-scanning SBS scheme24 or picosecond lasers in an impulsive SBS
arrangement29. Although these systems provide high spectrometer resolution, their low sensitivity
hinders biomechanical imaging.
To enable biomechanical SBS imaging, we constructed a SBS microscope (Fig. 1c; see Methods
Section 1 and Supplementary Fig. 1) based on our previous frequency-scanning SBS spectrometer
design26 -- 28, but with an improved spatial resolution of ~0.8 0.8 16 m3 and an enhanced shot-
noise sensitivity that yields a spectral acquisition time as low as 2 ms in water with Brillouin shift
and linewidth measurement precision of 11.5 MHz and 35 MHz, respectively (see Methods Section
2 and Supplementary Fig. 2). This acquisition time represents a 50-fold improvement over the
typical spectral measurement times of hundreds of milliseconds used in Brillouin measurements
of water11-13,16 -- 19. The obtainable Brillouin shift and linewidth measurement precision allow to
determine the real and imaginary parts of the gigahertz-frequency longitudinal modulus of water
to a fractional precision of 0.0046 and 0.11, respectively (see Methods Section 2). The light sources
were 780-nm tunable distributed-feedback semiconductor lasers that delivered a total excitation
power of ~265 mW at the sample. The linewidth of the lasers was 1-4 MHz, enabling highly
2
specific measurements of a SBG or SBL spectrum with megahertz spectrometer resolution, which
is at least 12-fold higher than the >100-MHz spectrometer resolution of Brillouin biomicroscopes9 --
20.
Fig. 1. Principle and method of SBS microscopy. (a) Acousto-optic interaction between focused, counter-propagating
pump (ω1) and probe (ω2 < ω1) light with longitudinal acoustic phonon (B). The interference of the pump and probe
fields (stripe pattern with intensity profile in white) drives the acoustic wave (gray wavefronts). (b) Intensity transfer
S) light by virtue of SBS along an interaction length l, with an intensity increase
between the pump (I1) and the probe (I2
in the probe beam (SBG) and an intensity decrease in the pump beam (SBL). Left and right illustrations are in the
frequency and spatial domains, respectively. (c) Schematic of the SBS microscope. The pump beam is modulated by
an acousto-optic modulator (AOM) and is left circularly polarized by a quarter wave plate (/4). The probe beam is
right circularly polarized by a second quarter wave plate (/4). The pump and the probe beams are focused to the same
point in the sample (S) by identical lenses (L). The frequency of the probe beam is scanned around the Stokes
frequency ω1 -- B to produce the backscattering SBG spectrum of the focus point on the probe intensity. The probe
beam is directed to an atomic notch filter (AF) by a polarizing beam splitter (PBS). The filter passes the probe light to
a custom transimpedance receiver (TIR) and suppresses the unwanted pump back-reflections into the receiver. The ac
signal at the receiver is input to a lock-in amplifier (LIA) to measure the SBG spectrum, whereas the dc signal is
processed by a personal computer (PC) to measure the total attenuation across the sample. By raster-scanning the
sample through the mutual focus of the pump and probe beams, three-dimensional images are obtained. (d) SBG
3
spectrum from volumes inside the pharynx (upper spectrum; purple cross in the coregistered brightfield image, scale
bar = 20 m) and the surrounding region (bottom spectrum; cyan cross in the coregistered brightfield image) of a live
wild-type C. elegans nematode. The spectra measured (T, dots) in 20 ms show a multipeak feature that was fitted well
by a double lorentzian line shape (T, black line), with contributions from a particular mechanical constituent of the
nematode (N, green line) and the buffer component (B, red line). (e) Three-dimensional sections of the Brillouin shift
B, Brillouin linewidth B, and peak SBG GB, through the head of a live C. elegans nematode. A pixel dwell time of
20 ms was used to acquire a SBG spectrum over the range of 3-7 GHz in each pixel. Images contain 100 200 pixels,
resulting in a total image recording time of 400 s.
We show in Fig. 1d the SBG spectrum from volumes inside the pharynx and the surrounding region
of a live wild-type C. elegans nematode (see Methods Sections 1 and 3, and Supplementary Fig.
3), which is an important model organism for biomedical research32. No substantial photodamage
was observed (see Methods Section 4 and Supplementary Fig. 4). A multipeak feature is apparent
in these two spectra, which can be well reproduced by a double lorentzian line shape. Whereas
the green lorentzian represents the SBG spectral contribution of a particular mechanical
constituent of the nematode, and the red lorentzian corresponds to the SBG spectral contribution
of the buffer component, which always exists in the sampled volume. Thus, SBS enables highly
specific identification of the Brillouin shift B, linewidth , and peak gain GB that completely
parametrize the SBG spectrum of a specific biomechanical component of the specimen. By raster
scanning the sample through the focus of the microscope, cross-sectional images of B, , and
GB of the nematode were obtained across the entire 50 m thickness of the worm's head, as
shown in Fig. 1e.
As the first application of SBS microscopy, we imaged and quantified the gigahertz-frequency
complex longitudinal modulus in live wild-type C. elegans at the second larval stage (L2) (Fig. 2).
Figure 2a displays the Brillouin shift B image and the coregistered brightfield image of a live C.
elegans larva along with images of the Brillouin linewidth (Fig. 2b) and peak gain GB (Fig.
2c) across the entire middle plane of the worm at a depth of 10 m. The pharynx, analogous to
the vertebrate esophagus, is an important organ of the nematode that pumps and grinds food into
the worm's intestine33. The pharynx is seen to exhibit high values of Brillouin shift and linewidth
but low values of peak gain compared to the surrounding region. From the SBG spectral
parameters (B, , GB) measured in the worm, and using the relation between mass density
and SBG23 and the reported refractive index distribution of the worm34, the spatial distributions
of the gigahertz-frequency complex longitudinal moduli in the nematode were evaluated, as shown
in Fig. 2d,e (see Methods Sections 5 and 6, and Supplementary Fig. 5). These distributions indicate
that at high frequencies the larva pharynx is stiffer -- with higher Brillouin shifts (Fig. 2a) -- and
more viscous -- with larger Brillouin linewidths (Fig. 2b) -- than its surroundings, attributable to
the muscular nature of the pharynx33. Such direct correlations between the Brillouin shift and
stiffness and the Brillouin linewidth and viscosity were consistently observed in other highly
hydrated biological systems10, 17. We further confirmed these findings by statistically comparing
the Brillouin shift and linewidth as well as the complex longitudinal modulus of the pharynx
(purple in Fig. 2a) and the surrounding region (cyan in Fig. 2a) averaged across the corresponding
volumes in several worms (Fig 2f-i). Therefore, we can use SBS microscopy to map the gigahertz-
frequency complex longitudinal modulus of live organisms, opening possibilities to study
biomechanics at high frequencies.
4
Fig. 2. Characterization of the high-frequency complex longitudinal modulus in live wild-type C. elegans L2 larvae
with SBS microscopy. (a) Brillouin shift B image (and the coregistered brightfield image) of a live L2 larva at a
depth of 10 m inside the worm. In the brightfield image, the pharynx (P) and the surrounding region (S) are filled in
purple and cyan, respectively. The pixel dwell time and the frequency acquisition range were as in Fig. 1e. Images
contain 100 400 pixels, resulting in a total image recording time of 800 s. Scale bar, 20 m. (b) Brillouin linewidth B
image of the larva. (c) Peak SBG GB image of the larva. (d) Gigahertz-frequency longitudinal storage modulus M' of
the larva (relative to that of double-distilled water M'W = 2.056 GPa). Mean refractive index values from the literature
were used (34) for the pharynx (n = 1.38) and the surrounding region (n = 1.36). For the other regions of the worm, a
mean refractive index value of 1.37 was used34. (e) Gigahertz-frequency longitudinal loss modulus M'' of the larva
(relative to that of double-distilled water M''W = 0.1249 GPa). (f) Average Brillouin shift of the pharynx (P, purple
bar) and the surroundings (S, cyan bar) of live L2 larvae (N = 10, p-value < 0.05 as computed by paired student's t-
test, 10-m depth). Error bars represent standard deviation from the mean. (g) Relative average gigahertz-frequency
longitudinal storage modulus M' of the larva pharynx and the surrounding region (same statistics as in f). (h) Average
Brillouin linewidth of the larva pharynx and surroundings (same statistics as in f). (i) Relative average gigahertz-
frequency longitudinal loss modulus M'' of the larva pharynx and surroundings (same statistics as in f). These results
indicate that at gigahertz frequencies the larva pharynx is stiffer and more viscous than its surroundings.
We also present the use of SBS microscopy to visualize variations in the high-frequency
viscoelastic response to hyperosmotic stress within the pharyngeal region of live wild-type C.
elegans young adults. Common instruments for biomechanics, such as cantilevers and
micropipettes, are limited to measurements at the surface of the worm and cannot image
biomechanical properties in vivo with depth sectioning2, 35 -- 38. Figure 3A-C shows the brightfield
images and the Brillouin shift B and linewidth B sections of the head region of three individual
C. elegans young adults from a depth of 25 m, under isotonic (Fig. 3a) and hypertonic (Fig. 3b,c)
salt conditions. At 125 mM of salt (250 Osm), the images show a notable increase in the Brillouin
shift and linewidth throughout the worm's head compared to a control worm under isotonic
conditions. At 250 mM of salt (500 Osm), the Brillouin shift and linewidth further increased.
These results suggest that hyperosmotic shock increases the high-frequency stiffness and
viscosity of the worm's pharyngeal region; consistent with previous findings that exposure of C.
elegans to a high-salt environment induces an increase in the low-frequency stiffness of the
animal's body-wall37. Similar results were reported in biological cells using spontaneous
Brillouin microscopy and atomic-force microscopy10. At both isotonic and hypertonic salt
concentrations, the Brillouin shift and linewidth of the worm's pharynx are found to remain
higher than those of the surrounding area, indicating that the pharynx is stiffer and more viscous
5
than its surroundings regardless of the osmotic conditions. All these findings were further verified
statistically in several nematodes by means of the average Brillouin shift and linewidth of the
pharynx and the surrounding region (Fig. 3d,e). These results show that SBS microscopy offers a
new approach for investigating biomechanics in vivo.
Fig. 3. SBS imaging of the variations in the high-frequency viscoelastic response to hyperosmotic stress within the
pharyngeal region of live wild-type C. elegans young adults. (a-c) Brillouin shift B and Brillouin linewidth B images
(and the coregistered brightfield image) of the head of live nematodes under isotonic salt conditions (see a), hypertonic
salt conditions (250 Osm, b), and hypertonic salt conditions (500 Osm, c). All the data was acquired from a depth of
25 m inside the worms' head. The arrow lengths in the worm schematic illustrate the rate of water entrance to and
exit from the nematode. The pixel dwell time and the frequency acquisition range were as in Fig. 1e. Images contain
100 200 pixels, resulting in a total image recording time of 400 s. Scale bar, 20 m. (d) Average Brillouin shift of
the pharynx (P, purple bar) and the surroundings (S, cyan bar) of live young adults (N = 10, p-value between levels of
osmolarity < 0.05 as computed by one-way ANOVA following post-hoc analysis Tukey's tests, p-value within levels
of osmolarity < 0.05, as computed by paired student's t-test, 25-m depth) before and after hyperosmotic shock of 125
mM NaCl (250 Osm) and 250 mM NaCl (500 Osm). Error bars represent standard deviation from the mean. (e)
Average Brillouin linewidth of the nematodes' pharynx and surroundings before and after the hyperosmotic shock
(same statistics as in d). These findings suggest that the gigahertz-frequency stiffness and viscosity of the worm's
pharyngeal region increases under hyperosmotic stress, whereas the pharynx is stiffer and more viscous than its
surroundings regardless of the osmotic conditions.
Another application of SBS microscopy is biomechanical mesoscopy. Although spontaneous
Brillouin microscopy has been used to acquire mesoscopic images of the Brillouin shift of tissue11,
mouse embryo15, and zebrafish18, the recording periods are often excessively long due to prolong
pixel dwell times (hundreds of milliseconds). While shorter times (50 ms) are possible for imaging
thin biological cells with spontaneous Brillouin microscopes based on virtually imaged phase
arrays20, their sub gigahertz spectrometer resolution considerably limits the ability to detect
biomechanical constituents with highly overlapping Brillouin bands17. We map here in vivo the
mesoscopic distributions of the Brillouin shift and the Brillouin linewidth across the middle plane
of C. elegans nematodes at three larval stages (L2, L3, L4) and two adult stages (young adult and
adult), as shown in Fig. 4a,b. These distribution maps allow us to identify organs and structures,
such as the pharynx, vulva, gonad, and eggs, within the worms during development based on their
viscoelastic contrast. Focusing on the nematode's head, a noticeable feature in these maps is the
larger Brillouin shifts and linewidths of the pharynx than those of its surroundings, at all
6
developmental stages. This was further validated statistically in several nematodes by means of
the average Brillouin shift and linewidth of the pharynx (inset of Fig. 4a) and the surrounding
region (inset of Fig. 4b). These findings imply that, regardless of the particular developmental
stage of the nematode, the pharynx is stiffer and more viscous than the surrounding area at high
frequencies. Interestingly, the mean Brillouin shift and linewidth of the pharynx and surroundings
remained relatively unchanged during development. This suggests that the mechanical properties
of the pharynx and the surrounding area are determined early in development, perhaps because of
the prominent function of the pharynx in feeding33.
Fig. 4. Biomechanical SBS mesoscopy of developing C. elegans nematodes. (a) Brillouin shift B images of live
nematodes at three larval stages (L2, L3, L4) and two adult stages, young adult (YA) and adult (A). The inset shows
the average Brillouin shift of the pharynx (P, purple bar) and the surroundings (S, cyan bar) of the live worms at the
different stages (N = 10 per stage, p-value within stages < 0.05 as computed by paired student's t-test, p-value for the
pharynx between the L4 stage and the other stages < 0.05, all other p-values between stages were > 0.05 as computed
by one-way ANOVA following post-hoc analysis Tukey's tests). Error bars represent standard deviation from the
mean. Images were acquired from the middle plane of the nematodes at depths of 10-35 m inside the worms. The
pixel dwell time and the frequency acquisition range were as in Fig. 1e. Images contain 100 400 -- 120 1200 pixels,
resulting in a total image recording time of 800 -- 2880 s. Scale bar, 50 m. (b) Brillouin linewidth B images of the
live nematodes at the different developmental stages. The average Brillouin linewidth of the worms' pharynx and
surroundings is displayed in the inset for the various stages (same statistics as in a). SBS microscopy enables in vivo
mechanical contrast visualization across large areas inside live organisms.
With its high sensitivity and spectrometer resolution, SBS microscopy enables mechanical
imaging of large and living samples in practical times and the ability to highly specify mechanical
constituents inside the samples based on resolvable Brillouin information. Together, these
7
capabilities may open up new possibilities for studying various mechanical aspects of cancer and
muscle disorders in model organisms.
Methods
1. SBS microscope
A detailed schematic of the SBS microscope is shown in Supplementary Fig. 1a. An amplified
continuous-wave (CW) distributed feedback (DFB) laser (SYST TA-pro-DFB, Toptica) and a CW
DFB laser (SYST DL-100-DFB, Toptica), both s-polarized, thermally stabilized and coupled into
a polarization-maintaining single-mode fiber, serve as a pump beam at a frequency 1 and a probe
beam at a frequency 2, respectively. The frequency 1 corresponds to the wavelength of 780.24
nm, and the frequency 2 is scanned around the Stokes frequency ω1 -- B, with B representing
the characteristic longitudinal Brillouin shift of the material. Although SBG and SBL spectra
provide equivalent information, we acquired SBG spectra because the passband of the detection
filter used is flatter for the Stokes than for the anti-Stokes frequency band26. The lasers exhibit
mode-hop-free tuning ranges larger than 50 GHz and laser linewidths of 1-4 MHz, providing wide
spectral range and high spectrometer resolution for the acquisition of SBG spectra. The frequency
of the probe laser is scanned by modulating its current with a sawtooth wave from a function
generator (AFG2021, Tektronix). The frequency difference between the pump and probe lasers
= 2 -- 1 is determined from a lookup table produced by measuring the beat frequency of the
two lasers as a function of the probe laser current with a fast photodetector (1434, Newport) and a
frequency counter (EIP 578B, Phase Matrix). The pump beam is optically filtered by a narrowband
Bragg filter (SPC-780, OptiGrate) to suppress the amplified spontaneous emission of the pump
laser. The filtered beam is modulated with an acousto-optic modulator (15210, Gooch and
Housego) driven by a 1.1 MHz sinusoidal wave from the function generator. The same sinusoidal
wave is used as a reference for the lock-in amplifier. Following modulation, the pump beam is p-
polarized by a half wave plate and is transmitted toward the sample through a polarizing beam
splitter. The pump and probe beams are circularly polarized orthogonally to each other with quarter
wave plates, and are tightly focused at a joint point in the sample by aspheric lenses with a
numerical aperture (NA) of 0.25 (Asphericon). This polarization scheme allows an efficient
modulation of the pump beam and also SBS in a backscattering geometry, which maximizes the
obtainable SBS signal. As a nonlinear process, SBS yielded with this NA an estimated lateral
resolution of / (4 NA) = 0.78 m and an estimated axial resolution of n / NA2 = 16.6 m
in water. The sample is mounted on a close-loop XYZ motorized stage (LS-50, Applied Scientific
Instrumentation) to enable raster scanning of the sample through the joint focus of the aspheric
lens pair. The backscattered SBG signal at the probe frequency 2 is collected with the same
aspheric lens that focuses pump light into the sample, is then s-polarized by the quarter wave plate
of the pump beam, and is finally reflected toward the detection module through the polarizing
beam splitter. The detection unit comprises an atomic vapor notch filter at the pump frequency 1,
an amplified photodetector, and a lock-in amplifier. The atomic filter is based on the narrowband
absorption of a rubidium-85 cell (SC-RB85-25x150-Q-AR, Photonics Technologies), heated to
90C, at the Fg=3 line frequency to which the pump laser is tuned for suppressing unwanted pump
stray reflections in the lock-in detection bandwidth. The amplified photodetector uses a large-area
photodiode (FDS1010, Thorlabs) reverse-biased at 24V and a novel transimpedance amplifier with
gain and equivalent input noise of 5104 V/A and 2.5 pA/Hz at the lock-in modulation frequency
8
of 1.1 MHz, respectively, to generate an ac-coupled output voltage proportional to the SBG. In
addition, the amplified photodetector has a dc-coupled voltage output to monitor the probe
intensity attenuated by transmission through the sample. A high‐frequency lock‐in amplifier
(SR844, Stanford Research) is used to demodulate the probe intensity. A data acquisition unit (NI
USB-6212 BNC, National Instruments) connected to a personal computer is fed with the analog
x-output of the lock-in amplifier, the dc-coupled voltage output of the amplified photodetector, the
sawtooth wave output of the function generator, and the analog readout channel of the frequency
counter.
To acquire a SBG spectrum, the lock-in detection bandwidth is set in the lock-in amplifier and the
sawtooth wave is programmed in the function generator, with the height of the waveform defining
the scanning frequency range of the probe laser around the Stokes frequency ω1 -- B and the time
period of the waveform determining the acquisition time of the entire spectrum. The microscope
is focused into a point in the sample and a custom program (LabVIEW version 2014, National
Instruments) tunes the frequency of the probe laser and records the dc and ac intensity of the
transmitted probe beam for each . Correction of the SBG signal for sample attenuation is
performed by estimating the attenuation coefficient of the sample and the pump intensity at the
focusing point from the measured dc probe intensity and the predetermined imaging depth in the
sample. Spectral analysis was performed in MATLAB (The MathWorks Inc.) to fit a measured
SBG spectrum to a double lorentzian line (Fig. 1d).
For biomechanical SBS imaging, the sample is raster-scanned through the microscope focus with
a XYZ motorized stage controlled by a multi-axis controller (MS-2000, Applied Scientific
Instrumentation). The beginning of a line scan is initiated by the controller, which also triggers,
via a transistor-transistor logic (TTL) signal, the function generator that drives the current of the
probe laser for frequency scanning (Supplementary Fig. 1b). The period of the driving current
defines the pixel dwell time over which a SBG spectrum is acquired with predetermined
measurement bandwidth and pixel size (Supplementary Fig. 1b). The pixel size (or the scanning
step size) should be at least two times smaller than the optical resolution for Nyquist sampling.
We set the pixel size to closely match the optical resolution (1 m), which represents a compromise
between the acquisition speed and the spatial resolution of the image. At each pixel, a SBG
spectrum is acquired by the lock‐in amplifier, as described above (Supplementary Fig. 1b). When
the motorized stage completes the line scan the TTL signal falls, which leads to termination of
frequency scanning and movement of the stage to the subsequent line scan, as shown in
Supplementary Fig. 1b.
For brightfield imaging, a halogen light source (HL-2000, Ocean Optics) was introduced into the
SBS microscope by placing a dichroic mirror in the path of the probe beam, and a complementary
metal -- oxide -- semiconductor camera (Lt225, Lumenera) was imaging the sample through a
dichroic mirror located immediately before the detection unit of the microscope, as shown in
Supplementary Fig. 1a.
2. Detection limit of the SBS microscope
We previously have demonstrated an SBS spectrometer for high-speed analysis of materials26 -- 28.
Here, we use a novel low-noise high-frequency transimpedance photoreceiver to yield shot-noise
limited sensitivity of the spectrometer. The electrical noise of our lock-in amplifier is 1.8 nV/Hz,
9
therefore dominating over the shot noise at detected optical powers of 0.1-10 mW. This noise
performance is insufficient for SBS bioimaging with high sensitivity. An efficient technique to
improve the sensitivity is to increase the signal-to-noise ratio (SNR) using a low-noise
transimpedance photoreceiver. To this end, a large-area photodiode (FDS1010, Thorlabs) was
integrated with a custom transimpedance amplifier with gain and equivalent input noise of 5104
V/A and 2.5 pA/Hz at the lock-in modulation frequency of 1.1 MHz, respectively. Using this
low-noise high-frequency transimpedance photoreceiver, the SBS microscope was brought close
to the shot-noise limit over the entire range of optical powers of 0.1-10 mW (Supplementary Fig.
2a). Further, we doubled the SBG signal detected by improving the crossing efficiency of the pump
and probe beams in the sample to 55% while minimizing the amount of pump stray reflections
detected. This was accomplished using measurement chambers that are much thicker than the
confocal parameter of the focusing lenses. Consequently, we were able to measure SBG spectra of
water in acquisition times as low as 2 ms with SNRs > 25 dB, resulting in Brillouin shift and
linewidth measurement precision of B/2 11.5 MHz and B/2 35 MHz, respectively
(Supplementary Fig. 2b-d). This result represents a fivefold improvement in the spectral
acquisition time compared with the previous SBS spectrometer, without compromising precision
of the measurements. For the mean Brillouin shift and linewidth retrieved from the SBG spectra
of water measured in 2 ms (B/2 = 5.05 GHz, B/2 = 323 MHz), the SBS microscope obtained
fractional precision of at least 2 B/B = 0.0046 and [(B/B)2 + (B/B)2]½ = 0.11 in the
estimate of the real and imaginary parts of the gigahertz-frequency longitudinal modulus M* of
water, respectively.
3. Preparation of the C. elegans samples
Wild type (N2) C. elegans worms were grown on nematode growth media (NGM) plates seeded
with the Escherichia coli OP50 or OP50-1 strains at 15°C. 30-60 embryos, laid at 15°C, were
picked, transferred to new plates and grown at 25°C for the duration of the experiment. Nematodes
at a well-defined developmental stage were determined using a light microscope during the
developmental time window of interest. During the reproductive period, animals were transferred
to fresh plates every 1-2 days to circumvent progeny contamination.
For imaging, we first prepared two 0.5-mm-thick 5% agar pads mixed with 10 mM sodium azide
solution (NaN3) to anesthetize the worms. Then, the agar pads were mounted on two 0.17-mm-
thick round glass coverslips (18 mm and 25 mm in diameter), and 10-15 nematodes at a specific
developmental stage were sandwiched between the agar-padded coverslips (Supplementary Fig.
3a). 10 l of M9 contact buffer was added between the agar pads and served as a control
environment for the worms. To fix the entire sample and to avoid dehydration, a few drops of UV
glue were applied at the edge of the smaller coverslip and a thin layer of Vaseline sealed the gap
between the two coverslips, as illustrated in Supplementary Fig. 3b. Following SBS imaging,
worms were washed from the agar pads to NGM plates for recovery.
4. Measurements of photodamage in SBS imaging of live C. elegans nematodes
In live-organism SBS imaging, high irradiation intensities are used. To assess photodamage, we
monitored the Brillouin shift at multiple locations in the head of live C. elegans nematodes and
across the agar pads surrounding the worm over 120 s at room temperature (20C) (Supplementary
Fig. 4). These measurements showed no consistent elevation in the Brillouin shift with time.
10
Because the Brillouin shift increases with rising temperature39, these results suggest that no
substantial sample heating occurred in 120 s − a time period significantly longer than the pixel
dwell time of 20 ms used in SBS imaging of live C. elegans worms. In addition, the maximum
standard deviation of these Brillouin shift measurements is 0.021 MHz, corresponding to a
maximum temperature change of less than 1.8C in water40. This result is consistent with our
calculation of the heating in water (below 1 K) due to water absorption at 780 nm using the same
conditions as in the experiments (total excitation power of 265 mW, NA of 0.25, and measurement
time of 120 s). In addition, examination of several worms following SBS imaging showed recovery
from anesthesia and a typical crawling motion.
5. Estimate of the mass density and the complex longitudinal modulus of materials by SBS
In SBS, the mean density of a material is related to the Brillouin shift B and linewidth B, and
the normalized line-center gain factor
of the SBG spectrum of the medium by the expression
,
(1)
where γe is the electrostrictive constant, ω2 is the frequency of the probe beam, n is the refractive
index of the medium, c is the speed of light in vacuum, and qB is the acoustic wavenumber.
is
calculated from the SBG spectrum G() as
,
(2)
with GB being the peak gain G(), the crossing efficiency of the pump and probe beams in the
sample, I1 the intensity of the pump beam at the sample entrance, and Leq the equivalent SBS
interaction length in the medium calculated using the assumed spatial distribution of the pump
intensity in the sample28. For backward SBS, qB = 2 ω2 n / c and Eq. (1) reduces to23
.
(3)
By estimating γe through use of the Lorentz -- Lorenz law, it is obtained that γe = (n2 -- 1) (n2 + 2)
/ 3 (Ref. 23). Thus, the mean density of a material can be evaluated from the measured values of
B, B, and GB of the SBG spectrum of the material and from knowledge about the material
refractive index n. Subsequently, the complex longitudinal modulus of the material M* is
calculated using the spectral parameters B and B of the measured SBG spectrum, the
estimated by Eq. (3), and the literature refractive index value of the material n as
,
(4)
11
NBg222B22BB2eNBqncg=NBgNBBeq1GgLI=2324NBBB2ecg=221BB2B*14Min=+with 1 denoting the wavelength of the beam at 1 (Refs. 7 and 30).
We validated Eq. (3) and computed M* (Eq. (4)) in various materials using the SBS microscope
with 0.033 NA focusing lenses (Supplementary Fig. 5). The use of these low NA lenses minimized
the effects of broadening and blue shift of the SBG spectrum and enabled to work in a near
backward SBS geometry with a crossing efficiency close to 100%. Reported refractive index
values at a wavelength of 780.24 nm and at 20C were used in Eqs. (3) and (4) for the materials
measured41,42. The estimated mass density of all materials agreed well, to within 1-10%, with
values from the literature (Ref. 43; Supplementary Table I).
6. Estimate of the mass density and the complex longitudinal modulus in C. elegans by the
SBS microscope
In backward SBS, the SBG spectrum of the ith voxel of volume
centered at
in a scattering medium is25
,
(5)
where the crossing efficiency of the pump and probe beams in the sample, g() is the SBS gain
factor which, to good approximation, is described by a lorentzian with a spectral shift ΩB, linewidth
B, and peak gain GB.
is the effective SBS interaction length in the medium
,
(6)
with
being the mean total attenuation coefficient of the medium at
.
is a
function that describes the dependence of the SBS interaction length on the spatial distribution of
the pump intensity in the sample28.
is the pump intensity at the voxel entrance
,
(7)
where I10 is the pump intensity at the entrance plane of the medium.
To estimate the mass density of the C. elegans nematode at the ith voxel, we used Eq. (3) with the
reported refractive index n value of the region to which the voxel belongs in the worm (34), the
Brillouin shift B and linewidth B retrieved from the SBG spectrum G(Ω) measured at the voxel,
and the normalized Brillouin line-center gain factor
evaluated as
.
(8)
Here, the mean total attenuation coefficient
, appearing in
(Eq. (6)) and
(Eq. (7)),
was recovered by measuring the attenuation of the probe beam through the worm at
using
12
xyz(),,iiixyz()()()1iieqeffGgfLI=ieffL()()(),1,1tiixyziefftiiLxye−−=−(),tiixy(),iixy()eqf1iI()(),2110tiiixyzziIIe−−=N,Big()()BN,B1iiieqeffGgfLI=(),tiixyieffL1iI(),iixythe dc output signal of the photoreceiver, together with estimating the thickness of the nematode
at the same point from the worm's brightfield images (assuming cylindrical symmetry of the
in Eqs. (6) and (7) was approximated as the extent of the
worm's body). The voxel's axial size
resolution cell along the axial dimension, and the voxel's axial position
in Eq. (7) was obtained
from the position readings of the motorized sample stage.
To compute the complex longitudinal modulus of the material M* at the ith voxel of the C. elegance
nematode, Eq. (4) is used with the spectral parameters B and B of the SBG spectrum measured
at the ith voxel, the mass of density estimated at the voxel, and the literature refractive index n
value of the region to which the voxel belongs in the worm.
We point out that Eq. (5) assumes low NA values (< ~0.1). For the NA of 0.25 used in SBS imaging
of the C. elegance worms, the Brillouin shift is expected to show a <1% decrease compared with
the Brillouin shift at NA = 0.1, the Brillouin linewidth a <5% increase, and the normalized
Brillouin line-center gain factor a <4% decrease, resulting in a <2% (<4%) decrease (increase) in
the calculation of the real (imaginary) part of the complex longitudinal modulus.
References
1. Thomas, G., Burnham, N. A., Camesano, T. A. & Wen, Q. Measuring the mechanical properties of living cells
using atomic force microscopy. J. Vis. Exp. 76, 50497 (2013).
2. Essmann, C. L. et al. In-vivo high resolution AFM topographic imaging of Caenorhabditis elegans reveals
previously unreported surface structures of cuticle mutants. Nanomedicine 13, 183-189 (2017).
3. Goulam Houssen, Y., Gusachenko, I., Schanne-Klein, M. C. & Allain, J. M. Monitoring micrometer-scale
collagen organization in rat-tail tendon upon mechanical strain using second harmonic microscopy. J. Biomech.
44, 2047 -- 2052 (2011).
4. Wang, S. & Larin, K. V. Shear wave imaging optical coherence tomography (SWI-OCT) for ocular tissue
biomechanics. Opt. Lett. 39, 41-44 (2014).
5. Kennedy, K. M. et al. Quantitative micro-elastography: imaging of tissue elasticity using compression optical
coherence elastography. Sci. Rep. 5, 15538 (2015).
6. Kennedy, B. F., Wijesinghe, P. & Sampson, D. D. The emergence of optical elastography in biomedicine. Nat.
Photonics 11, 215 -- 221 (2017).
7. Vaughan, J. M. & Randall, J. T. Brillouin scattering, density and elastic properties of the lens and cornea of the
eye. Nature 284, 489-491 (1980).
8. Koski, K. J., Akhenblit, P., McKiernan, K. & Yarger, J. L. Non-invasive determination of the complete elastic
moduli of spider silks. Nat. Mater. 12, 262-267 (2013).
9. Palombo, F. et al. Biomechanics of fibrous proteins of the extracellular matrix studied by Brillouin scattering, J.
R. Soc. Interface 11, 20140739 (2014).
10. Scarcelli, G. et al. Noncontact three-dimensional mapping of intracellular hydromechanical properties by
Brillouin microscopy. Nat. Methods 12, 1132-1134 (2015).
11. Antonacci, G. et al. Quantification of plaque stiffness by Brillouin microscopy in experimental thin cap
fibroatheroma. J. R. Soc. Interface 12, 20150483 (2015).
12. Lepert, G., Gouveia, R. M., Connonb, C. J. & Paterson, C. Assessing corneal biomechanics with Brillouin spectro-
microscopy. Faraday Discuss. 187, 415-428 (2016).
13. Elsayad, K. et al. Mapping the subcellular mechanical properties of live cells in tissues with fluorescence
emission-Brillouin imaging. Sci. Signal. 9, rs5 (2016).
14. Meng, Z., Traverso, A. J., Ballmann, C. W., Troyanova-Wood, M. A. & Yakovlev, V. V. Seeing cells in a new
light: a renaissance of Brillouin spectroscopy. Adv. Opt. Photonics 8, 300-327 (2016).
15. Raghunathan, R. et al. Evaluating biomechanical properties of murine embryos using Brillouin microscopy and
optical coherence tomography. J. Biomed. Opt. 22, 086013 (2017).
13
ziz16. Karampatzakis, A. et al. Probing the internal micromechanical properties of Pseudomonas aeruginosa biofilms
by Brillouin imaging. NPJ Biofilms Micro. 3, 20 (2017).
17. Mattana, S. et al. Non-contact mechanical and chemical analysis of single living cells by microspectroscopic
techniques. Light Sci. Appl. 7, 17139 (2018).
18. Schlussler, R. et al. Mechanical mapping of spinal cord growth and repair in living zebrafish larvae by Brillouin
imaging. Biophys. J. 115, 911-923 (2018).
19. Bevilacqua, C., Sánchez-Iranzo, H., Richter, D., Diz-Muñoz, A. & Prevedel, R. Imaging mechanical properties
of sub-micron ECM in live zebrafish using Brillouin microscopy. Biomed. Opt. Express 10, 1420-1431 (2019).
20. Nikolić, M. & Scarcelli, G., Long-term Brillouin imaging of live cells with reduced absorption-mediated damage
at 660nm wavelength. Biomed. Opt. Express 10, 1567-1580 (2019).
21. Chiao, R. Y., Townes, C. H. & Stoicheff, B. P. Stimulated Brillouin scattering and coherent generation of intense
hypersonic waves. Phys. Rev. Lett. 12, 592-595 (1964).
22. Damzen, M. j. Stimulated Brillouin Scattering: Fundamentals and Applications (IOP Publishing, Bristol, 2003).
23. Boyd, R. W. Nonlinear Optics (Academic Press, New York, ed. 3, 2008).
24. Ballmann, C. W. et al. Stimulated Brillouin scattering microscopic imaging. Sci. Rep. 5, 18139 (2015).
25. Remer, I. & Bilenca, A. Background-free Brillouin spectroscopy in scattering media at 780 nm via stimulated
Brillouin scattering. Opt. Lett. 41, 926-929 (2016)
26. Remer, I. & Bilenca, A. High-speed stimulated Brillouin scattering spectroscopy at 780 nm. APL Photonics 1,
061301 (2016).
27. Remer, I., Cohen, L. & Bilenca, A. High-speed continuous-wave stimulated Brillouin scattering spectrometer for
material analysis. J. Vis. Exp. 127, e55527 (2017).
28. Remer, I. Stimulated Brillouin Scattering Microscopy. (Ben-Gurion University of the Negev, 2017).
29. Ballmann, C. W., Meng, Z., Traverso, A. J., Scully, M. O. & Yakovlev, V. V., Impulsive Brillouin microscopy.
Optica. 4, 124-128 (2017).
30. Litovitz, T. A. & Davis, C. M. Physical Acoustics, (Academic, New York, 1965, Vol. 2A)
31. Freudiger, C. W. et al. Label-free biomedical imaging with high sensitivity by stimulated Raman scattering
microscopy. Science 322, 1857-1861 (2008).
32. Kaletta, T. & Hengartner, M. O. Finding function in novel targets: C. elegans as a model organism. Nat. Rev.
Drug Discov. 5, 387-399 (2006).
33. Mango, S. E. The C. elegans pharynx: a model for organogenesis. WormBook, 1-26 (2007).
10.1895/wormbook.1.129.1.
34. Choi, W. et al. Tomographic phase microscopy. Nat. Methods 4, 717 -- 719 (2007).
35. Park, S. J., Goodman, M. B. & Pruitt, B. L. Analysis of nematode mechanics by piezoresistive displacement clam.
Proc. Natl. Acad. Sci. U S A. 104, 17376-17381 (2007).
36. Sznitman, J., Purohit, P. K., Krajacic, P., Lamitina, T. & Arratia, P. E. Material properties of Caenorhabditis
elegans swimming at low Reynolds number. Biophys. J. 98, 617-26 (2010).
37. Backholm, M., Ryu, W. S. & Dalnoki-Veress, K. Viscoelastic properties of the nematode Caenorhabditis elegans,
a self-similar, shear-thinning worm. Proc. Natl. Acad. Sci U S A. 110, 4528- 4533 (2013).
38. Gilpin, W., Uppaluri, S. & Brangwynne, C. P. Worms under pressure: bulk mechanical properties of C. elegans
are Independent of the cuticle. Biophys J. 108, 1887-1898 (2015).
39. Fry, E. S., Emery, Y., Quan, X. & Katz, J. W. Accuracy limitations on Brillouin lidar measurements of
temperature and sound speed in the ocean, Appl. Opt. 36, 6887-6894 (1997).
40. Schonle, A. & Hell, S. W. Heating by absorption in the focus of an objective lens. Opt. Lett. 23, 325-327 (1998).
41. Podeǎ, J., Procházka, O. & Medin, A. Studies on agaroses; 1. Specific refractive index increments in dimethyl
sulfoxide and in water at various wavelengths and temperatures. Polymer 36, 4967-4970 (1995).
42. Polyanskiy, M. N. Refractive index database. Available at: http://www.refractiveindex.info (2008-2019).
43. Weast, R. C. Handbook of Chemistry and Physics (CRC Press, 1987).
14
Supplementary Information
High Sensitivity and Specificity Biomechanical Imaging by
Stimulated Brillouin Scattering Microscopy
Itay Remer1,2, Netta Shemsesh3, Anat Ben-Zvi3 & Alberto Bilenca1,4*
1Biomedical Engineering Department, Ben-Gurion University of the Negev, 1 Ben Gurion Blvd, Be'er-Sheva 84105,
Israel
2Agilent Research Laboratories, 94 Shlomo Shmeltzer Road, Petach Tikva 4970602, Israel
3Department of Life Sciences, Ben-Gurion University of the Negev, 1 Ben Gurion Blvd, Be'er Sheva 84105, Israel
4Ilse Katz Institute for Nanoscale Science and Technology, Ben-Gurion University of the Negev,
1 Ben Gurion Blvd, Be'er-Sheva 84105, Israel
e-mail: *[email protected]
15
Supplementary Fig. 1. Detailed schematic of the SBS microscope. (a) Near infrared (NIR) fiber-coupled pump and
probe distributed feedback (DFB) lasers are collimated with collimators of 1 mm (C). The collimated, s-polarized
probe laser beam (light red) is expanded by lenses L1 and L2, and is then right circularly polarized by a quarter wave
plate (/4) and focused into the sample (S) by a 0.25 NA aspheric lens L3. The collimated, s-polarized pump laser
beam (deep red) is filtered with a Bragg filter (BF) and is then expanded by lenses L4 and L5 and sinusoidally
modulated by an acousto-optic modulator (AOM) at 1.1 MHz. Next, the beam is p-polarized by a half wave plate (/2)
and transmitted through a polarizing beam splitter (PBS). The beam is subsequently left circularly polarized by a
16
quarter wave plate (/4) and focused by an aspheric lens L6 (same as lens L3) to the focal point of lens L3 in the sample.
The backscattered SBG signal that modulates the intensity of the probe beam is collected with lens L6 and is then s-
polarized by a quarter wave plate (/4) and reflected by the PBS. Next, the signal passes a rubidium notch filter (85Rb)
at the pump frequency, and is detected by a transimpedance receiver (TIR), which comprises a large area photodiode
(PD), a bias tee, and a transimpedance amplifier (TIA). The SBG signal in the ac output voltage of the TIR is measured
by a lock-in amplifier (LIA). Two fiber splitters (FS) extract auxiliary beams from the pump and probe lasers, and a
fast photodetector (FPD) and a frequency counter (FC) measure the beat frequency (or the difference frequency)
between these beams. The SBG signal, the average probe power (measured by the dc output voltage of the TIR), the
difference frequency measurement, and the probe laser current Iprobe are all sampled by a data acquisition card (DAQ)
connected to a personal computer (PC) for further analysis and visualization. A wide halogen illumination and a
CMOS camera co-register brightfield images in the microscope. The halogen beam is directed from a collimated fiber-
coupled source to the sample through lenses L7 and L3 and a dichroic mirror (DM) that combines white and NIR light.
The sample is imaged onto the camera via lenses L6 and L8, where a DM positioned after the PBS splits the NIR and
white light for SBG and brightfield imaging, respectively. The camera is connected to a PC for further analysis and
presentation. All folding mirrors (M) fit the microscope on a 18''×24'' breadboard that is vertically mounted on an
optical table. (b) SBS imaging acquisition scheme. The sample stage controller initiates the beginning of a line scan
and TTL-triggers the probe current Iprobe to start the frequency scans. The motorized stage raster scans the sample
through the joint focus of lenses L3 and L6, while the probe laser scans the pump-probe frequency difference around
the Brillouin shift of the sample and the LIA retrieves the SBG spectrum at a pixel. The TTL signal falls when the
stage completes the line scan and this terminates the frequency scans and moves the stage to the next line scan.
17
Supplementary Fig. 2. Performance of the SBS microscope. (a) Noise density against detected power. Measurements
(open circles) were made in double-distilled water at room temperature (20C) with only probe light excitation. To
measure the noise density, the probe frequency ω2 was scanned around the Stokes frequency ω1 -- B (~ −5.05 GHz
from the pump frequency ω1) over a 2-GHz bandwidth in 100 ms and the ac signal at the receiver was lock-in detected
at a 1.1-MHz reference frequency. The noise density was calculated as the ratio of the standard deviation of the ac
voltage to the square-root of the lock-in bandwidth (B = 138 Hz). The optical probe power was measured from the dc
voltage at the receiver. The fit of the noise density measurements to the sum of three noise contributions is shown in
a solid line. These noise contributions are the electrical noise density (125 nV/Hz), the shot noise density (2q R
P where q is the electron's charge, R = 0.53 A/W is the photodetector responsivity, and P is the optical power of the
probe beam) and the relative intensity noise of the laser (RIN R2 P2 with RIN = 3.16 10-17 Hz-1). (b) SBG
spectrum of double-distilled water at room temperature (20C). The spectrum (close circles) was acquired over a 2-
GHz bandwidth in 2 ms. A total excitation power of 265 mW was used. The lorentzian fit to the measurements is
shown in a red solid line. (c) Signal-to-noise ratio (SNR) against the acquisition time of the SBG spectrum of water.
The SNR was evaluated as the ratio of the SBG peak power to the standard deviation of the noise skirt in the SBG
spectrum over a 1-GHz bandwidth. Open circles and error bars represent mean value and standard deviation from the
mean, respectively, as calculated from one hundred SBG spectra measured sequentially. A slope of ~10 dB/dec was
obtained from the linear regression line, as expected from the direct dependence of the SNR on the acquisition time
(25). (d) Precision of the Brillouin shift and linewidth measurements (B and B) against the acquisition time of
the SBG spectrum of water. The precision was calculated as the standard deviation of the Brillouin shift and linewidth
recovered from the lorentzian fits of one hundred SBG spectra measured sequentially. Dashed lines are drawn to guide
the eye.
18
Supplementary Fig. 3. Sample mount for SBS imaging of live C. elegans nematodes. (a) Three-dimensional
schematic and side view of the sample mount. Nematodes are sandwiched between 0.5-mm-thick agar pads each is
mounted on a 0.17-mm-thick round glass coverslip of a different diameter (18 mm and 25 mm). (b) Top view of the
sample mount. A few drops of UV glue are applied at the edge of the smaller coverslip and a thin layer of Vaseline
seals the gap between the two coverslips, thereby fixing the entire sample and avoiding dehydration.
19
Supplementary Fig. 4. Photodamage in SBS imaging of live C. elegans nematodes. Brillouin shift B versus time
for a live C. elegans worm. The Brillouin shift B was recovered from a double lorentzian fit of the SBG spectra
measured over a 4-GHz bandwidth in 200 ms. For these measurements, a total excitation power of 265 mW was used,
which is the same power level employed all through this work. Curve colors represent different spatial positions across
the middle plane of the sample at which the measurements were taken (~25 m in depth). These positions are marked
in the coregistered brightfield image with crosses of corresponding colors (scale bar, 20 m). Curve pale and dark
shades indicate raw and ten-sample moving average data, respectively.
20
Supplementary Fig. 5. Estimation of the mass density and the complex longitudinal modulus of materials by SBS.
(a) Brillouin shift B of methanol, dimethyl sulfoxide (DMSO), double-distilled water, and 1%, 5% and 10%
agar/water gels at room temperature (20C). (b) Brillouin linewidth B of the different materials. (c) Normalized
Brillouin line-center gain factor
of the different materials. The spectral parameters (B, ,
) were retrieved
from the lorentzian fits of two hundred SBG spectra measured sequentially. Each spectrum was recorded over a
bandwidth of 4 GHz in 200 ms by the SBS microscope with 0.033 NA focusing lenses and a total excitation power of
285 mW. (d) Mass density estimate for the different materials using Eq. (3) of Methods Section 5 with refractive index
values of 1.315, 1.47, 1.329, 1.331, 1.337, and 1.345 for methanol, DMSO, double-distilled water, 1%, 5% and 10%
agar/water gels, respectively. (e) Real part of the gigahertz-frequency longitudinal modulus M' for the different
materials. (f) Imaginary part of the gigahertz-frequency longitudinal modulus M'' for the different materials. M' and
M'' were calculated using Eq. (4) of Methods Section 5 with the Brillouin shift B and linewidth B values described
in a and b, and the mass density estimates and refractive indices presented in d.
21
NBgNBgSupplementary Table I. Mean and standard deviation values for the Brillouin shift B, the Brillouin linewidth B,
the normalized Brillouin line-center gain factor
, the mass density estimate
, and the real and imaginary parts
of the gigahertz-frequency longitudinal modulus M' and M'' of methanol, DMSO, double-distilled water, and 1%, 5%
and 10% agar/water gels at room temperature (20C). Literature values for the mass density are also presented43.
22
NBg |
1701.09173 | 1 | 1701 | 2017-01-31T18:39:18 | Asymmetry of short-term control of spatio-temporal gait parameters during treadmill walking | [
"physics.bio-ph"
] | Optimization of energy cost determines average values of spatio-temporal gait parameters such as step duration, step length or step speed. However, during walking, humans need to adapt these parameters at every step to respond to exogenous and/or endogenic perturbations. While some neurological mechanisms that trigger these responses are known, our understanding of the fundamental principles governing step-by-step adaptation remains elusive. We determined the gait parameters of 20 healthy subjects with right-foot preference during treadmill walking at speeds of 1.1, 1.4 and 1.7 m/s. We found that when the value of the gait parameter was conspicuously greater (smaller) than the mean value, it was either followed immediately by a smaller (greater) value of the contralateral leg (interleg control), or the deviation from the mean value decreased during the next movement of ipsilateral leg (intraleg control). The selection of step duration and the selection of step length during such transient control events were performed in unique ways. We quantified the symmetry of short-term control of gait parameters and observed the significant dominance of the right leg in short-term control of all three parameters at higher speeds (1.4 and 1.7 m/s). | physics.bio-ph | physics |
Asymmetry of short-term control of spatio-temporal
gait parameters during treadmill walking
Klaudia Kozlowska1,+, Miroslaw Latka1,+*, and Bruce J. West2,+
1Wroclaw University of Science and Technology, Faculty of Fundamental Problems of Technology, Department of
Biomedical Engineering, Wroclaw, 50-370, Poland
2Army Research Office, Information Sciences Directorate, Research Triangle Park, 27709, USA
*[email protected]
+each author contributed equally to this work
ABSTRACT
Optimization of energy cost determines average values of spatio-temporal gait parameters such as step duration, step length
or step speed. However, during walking, humans need to adapt these parameters at every step to respond to exogenous
and/or endogenic perturbations. While some neurological mechanisms that trigger these responses are known, our under-
standing of the fundamental principles governing step-by-step adaptation remains elusive. We determined the gait parameters
of 20 healthy subjects with right-foot preference during treadmill walking at speeds of 1.1, 1.4 and 1.7 m/s. We found that
when the value of the gait parameter was conspicuously greater (smaller) than the mean value, it was either followed imme-
diately by a smaller (greater) value of the contralateral leg (interleg control), or the deviation from the mean value decreased
during the next movement of ipsilateral leg (intraleg control). The selection of step duration and the selection of step length
during such transient control events were performed in unique ways. We quantified the symmetry of short-term control of
gait parameters and observed the significant dominance of the right leg in short-term control of all three parameters at higher
speeds (1.4 and 1.7 m/s).
Introduction
It has been known for over a century that the stride interval of human gait is remarkably stable. Small fluctuations of approxi-
mately 3 -- 4% were attributed to the complexity of the locomotor system and treated as an uncorrelated random process.1 From
this viewpoint, the discovery of long-time, fractal correlations in stride-interval time series was unexpected.2, 3 Those early
papers not only spurred interest in the emerging concept of fractal physiology,4 but also shifted the focus of quantitative gait
analysis from average values of typical parameters (e.g. stride intervals) to their temporal variability. That profound change
of perspective brought new insights into locomotor manifestations of Huntington's and Parkinson's diseases, aging, and the
connection between gait dynamics and fall risk, see5 and references therein.
From a plethora of physiologically accessible gait patterns, humans employ only walking and running. Walking feels
easiest at low speeds, and running feels easiest when moving faster. Optimization of energy cost underlies not only the choice
of gait,6 -- 8 but also determines average values of gait parameters, such as step length and duration.10 During walking, humans
need to adapt their spatio-temporal gait parameters at every step to be able to respond to exogenous (e.g. irregularities of
walking surface) and/or endogenic (neuromuscular noise) perturbations.11 While some neurological mechanisms that trigger
these responses are known,12 -- 15 the fundamental principles governing step-by-step adaptation remain elusive.16
Treadmill walking, especially at high speeds, presents challenges that can be met only through effective short-term control
of spatio-temporal gait parameters. In order to stay on a treadmill, the subject's step duration and length must yield a step
speed which can fluctuate over a narrow range centered on the treadmill belt's speed. The results of previous experiments with
walking on a split-belt treadmill underscore the intricacies of such aggregation. In particular, spatial and temporal controls of
locomotion are accessible through distinct neural circuits17 and neural control of intra- versus interlimb parameters (calculated
using values from both legs, e.g., step length, double support) during walking is to a large extent independent.18 Herein,
we investigate the dynamics of time series of gait parameters (step duration, length and speed) following a sudden, large
deviation from the mean value. In particular, we test the hypothesis that whenever the value of a gait parameter is markedly
greater (smaller) than the mean value, it is either immediately followed by a smaller (greater) value of the contralateral
leg (interleg control), or the deviation from the mean value decreases during the next movement of ipsilateral leg (intraleg
control). Said differently, during treadmill walking errors are not gradually attenuated via long-term corrections, but are
corrected immediately by the same or opposite leg. Taking into account differences in the relative contribution of lower limbs
Table 1. The statistics of the occurrence of intraleg (columns L-L and R-R) and interleg (columns R-L and L-R) control
mechanisms that are evoked in response to errors in step duration, length and velocity. Statistics are presented for three
values of treadmill speed v. The number of errors, defined as abrupt deviations from the moving average value, is presented
in the Total column. The column labeled NC gives the number of errors that persisted for more than two successive steps.
The Left and Right columns show the number of control events performed by each leg. Probabilities of occurrence of intra-
and interleg control mechanisms are listed in the last four columns.
v [m/s] Total NC L-L R-L R-R L-R Left [%] Right [%]
pLL
pRL
pRR
pLR
1.1
1.4
1.7
1.1
1.4
1.7
1.1
1.4
1.7
180
121
103
380
257
186
393
275
189
4
11
5
14
12
18
5
14
8
42
29
22
125
73
42
51
21
18
71
30
17
83
63
34
115
61
45
32
18
26
122
93
87
84
61
66
step duration
51
42
32
53
45
41
step length
54
48
42
49
22
35
step speed
51
48
44
106
76
36
47
55
59
46
58
65
49
52
56
0.23
0.24
0.21
0.28
0.17
0.17
0.18
0.15
0.25
0.28
0.35
0.31
0.33
0.28
0.23
0.19
0.12
0.09
0.32
0.36
0.47
0.13
0.19
0.12
0.21
0.23
0.18
0.29
0.22
0.24
0.21
0.22
0.35
0.27
0.28
0.19
to control and propulsion -- the effect known as functional gait asymmetry,19 we further hypothesize that in subjects with right
foot preference the short-term control of gait spatio-temporal parameters is stronger for the right leg.
Results
In Table 1 we recorded data which are crucial for testing the main hypothesis of the paper: that errors in gait parameters are
e corrected immediately by the same or opposite leg. Let us focus on the first row of this table which concerns step duration
control at a treadmill speed of 1.1 m/s. There were 180 "errors" defined as abrupt changes in step duration (equations 2-4).
For the left leg, in 42 cases (column L-L), the deviation of step duration from the mean value did not trigger a compensating
change in the step duration of the right leg. The value of the control parameter DLR, defined by equation 5, greater than
1 indicates the absence of such adjustment. However, the deviation decreased during the next left step as indicated by the
value of intraleg control parameter DLL, defined by equation 7, smaller than 1 (the statistics of both inter- and intraleg control
parameters are presented in Table 2). In other words, for the left leg, we observed intraleg control of step duration in 42 cases.
In 51 cases (column R-L), the change in the step duration of the left leg compensated the deviation of step duration of the
previous right step (interleg control). By adding columns L-L and R-L, we obtain 93 control events performed by the left leg.
This number expressed as the percentage of all 180 control events is given in the column of Table 1 labeled as Left. Please
note that only in 4 (column NC) out of 180 cases (2.2%), the appearance of a step duration error did not evoke either of the
control mechanisms.
For all three gait parameters, the number of errors decreases with speed. For example, there were 180 errors in step
duration at v = 1.1 m/s but only 103 at v = 1.7 m/s (a 43% reduction). A comparable drop in the number of errors was
observed for step length (51%) and velocity (52%). Less than half of these changes can be explained by the 22% reduction
of the number of steps taken by all the subjects at v = 1.7 m/s in comparison with v = 1.1 m/s. Please note that the number
of steps decreases with treadmill speed since at each speed, the subjects were asked to cover the same distance of 400 m. It
is worth emphasizing that there were approximately twice as many errors in step length and speed than in step duration. For
all three parameters: step duration, length, and speed, in at least 90% of cases, the deviations from the mean value decreased
during the subsequent two steps via either intra- or interleg control.
Regardless of treadmill belt speed, the control of step length is predominately intraleg (Table 1). For example, for the
left leg at the lowest speed, the probability of evoking intraleg control pLL = 0.33 is 42% greater than that of interleg control
pRL = 0.19). In the same condition, for the right leg such difference is equal to 59% ( pRR = 0.32 vs pLR = 0.13). There is no
such pattern for the other two gait parameters.
2/11
Table 2. The values of intra- and interleg control parameters for treadmill walking. Data are presented as mean (standard
deviation).
v [m/s]
DLL
DRL
DRR
DLR
DBEL DBER
D DBE [%]
step duration
0.34 (0.22)
0.45 (0.26)
0.31 (0.25)
0.55 (0.26)
0.55 (0.25)
0.63 (0.20)
0.43 (0.25)
0.31 (0.22)
0.36 (0.28)
0.58 (0.21)
0.54 (0.24)
0.49 (0.26)
step length
0.46 (0.26)
0.46 (0.27)
0.44 (0.26)
0.72 (0.23)
0.79 (0.16)
0.79 (0.15)
0.48 (0.25)
0.45 (0.26)
0.44 (0.27)
0.62 (0.30)
0.61 (0.30)
0.68 (0.23)
step speed
0.44 (0.26)
0.39 (0.25)
0.51 (0.26)
0.65 (0.26)
0.68 (0.24)
0.73 (0.20)
0.40 (0.27)
0.33 (0.23)
0.41 (0.24)
0.56 (0.28)
0.56 (0.29)
0.58 (0.27)
1.20
0.85
0.97
0.98
0.76
0.63
0.92
0.91
0.68
0.91
1.12
1.34
0.88
1.12
1.24
1.01
1.15
1.19
1.1
1.4
1.7
1.1
1.4
1.7
1.1
1.4
1.7
28
-28
-32
11
-38
-65
-9
-23
-55
For step duration, length, and speed, the control parameter D was independent of speed (Table 2). For all three gait
parameters, both for the right and left leg, the mean value of D for interleg control was greater than that of intraleg control.
For example, for step duration at v = 1.1m/s DRL = 0.55 and DLL = 0.34.
The difference between the values of intra- and interleg control parameters for a given leg was statistically significant for
all three treadmill speeds for step length:
• at 1.1 m/s: ple f t < 1 × 10−4, pright = 6 × 10−3;
• at 1.4 m/s: ple f t < 1 × 10−4, pright = 3 × 10−2;
• at 1.7 m/s: ple f t = 1 × 10−4, pright = 2 × 10−3;
as well as the step speed:
• at 1.1 m/s: ple f t < 1 × 10−4, pright = 6 × 10−4;
• at 1.4 m/s: ple f t < 1 × 10−4, pright < 1 × 10−4;
• at 1.7 m/s: ple f t = 2 × 10−3, pright = 9 × 10−3.
For step duration such differences were not so strongly pronounced:
• at 1.1 m/s: ple f t = 5 × 10−4, pright = 5 × 10−2;
• at 1.7 m/s: ple f t = 2 × 10−3.
With the exception of step duration and step length at the lowest speed (v = 1.1 m/s), the asymmetry parameter D DBE was
smaller than zero indicating a dominant role of the right leg in short-term control of gait parameters during treadmill walking.
Table 3 shows the probability of compensatory response to errors in gait spatio-temporal parameters for intra- (L-L, R-
R) and interleg (L-R, R-L) control. Such response corresponds to negative values of variables Sinter (equation 6) and Sintra
(equation 8). For all speeds and parameters, the probability of interleg compensation is close to 1, roughly two times higher
than that of intraleg response.
3/11
Table 3. Probability of compensation of errors in gait spatio-temporal parameters for intra- (L-L, R-R) and interleg (L-R,
R-L) control. Statistics are presented for three values of treadmill speed v.
v [m/s]
L-L R-R L-R R-L
L-L
R-R L-R R-L
L-L R-R L-R R-L
step duration
0.98
0.42
0.95
0.39
0.54
0.97
0.50
0.28
0.50
0.94
0.95
0.94
0.40
0.42
0.40
step length
0.83
0.38
0.82
0.34
0.43
0.95
1.1
1.4
1.7
0.83
0.80
0.76
0.52
0.52
0.32
step speed
0.92
0.55
0.93
0.46
0.44
0.95
0.92
0.92
0.89
Discussion
In overground walking with self-selected speed, fluctuations of stride interval, length, and speed exhibit persistent fractal
scaling characterized by a Hurst exponent a > 0.5.2, 3, 20 Auditory metronomic cueing changes fractal statistics of stride
intervals from persistent to antipersistent (a < 0.5).21 The super central pattern generator model, introduced by West and
Scafeta,22 elucidates the dynamic origin of such transitions. In particular, the transitions result from the driving of a fractal
clock, which retains its properties under perturbation. In treadmill walking, fluctuations of interstride interval and stride length
are also persistent. However, the time series of stride speed is antipersistent, which is a manifestion of increased central control
of this gait parameter.16, 23 Terrier has recently demonstrated that visual cueing (alignment of step lengths with marks on the
floor) also induced anti-correlated pattern in gait parameters.30
To a large extent, fluctuations of spatio-temporal gait parameters result from the intrinsic fractal properties of pattern gen-
erators. Hidden in these fluctuations are sporadic control events, triggered to accomplish a locomotor task such as remaining
on a moving treadmill belt. This is why we study the dynamics of time series of gait parameters that follow a sudden large
deviation from a mean value. For lack of a better word, we dubbed such events errors, but emphasize that they may originate
either from the failure of the motor control system, or from the necessary adjustment of the subject's position on a treadmill.
While the definition of such events is arbitrary (equations 2-4), it satisfies the research objective.
We found that when the value of the gait parameter (step duration, length or speed) was conspicuously greater (smaller)
than the mean value, it was either followed immediately by a smaller (greater) value of the contralateral leg (interleg control),
or the deviation from the mean value decreased during the next movement of ipsilateral leg (intraleg control). The existence
of distinct short-term control of step frequency (the inverse of step duration) was demonstrated by Snaterse et al.24 The
time evolution of step frequency triggered by sudden stepwise increments in treadmill speed was modeled by the sum of two
exponentially decaying terms. The time constant of the first term was 1.44 ± 1.14 s and its amplitude was two times larger
than that of the second term, whose time constant was 27.56 ± 16.18s. For those values of time constants, step frequency
adjustments were two-thirds complete in less than two seconds. Snaterse et al. argued that the first term represents a rapid pre-
programmed response, while the slower one models fine-tuning of step frequency driven by energy expenditure optimization.
Herein we extended this line of reasoning by demonstrating that short-term control of gait parameters may be realized using
intra- and interleg adjustments. The better understanding of short-term control mechanisms does not bring us any closer to
understanding how, during treadmill walking, persistent stochastic variables: step duration and step length are combined to
yield antipersistent step speed. We believe that a different mechanism operating at a longer time scale underlies this effect.
There are fundamental differences between the control of step duration and step length. The probability of evoking intraleg
control of step length at the highest treadmill speed ( v = 1.7m/s) is approximately three times greater than that of evoking
interleg control. There is no such distinct pattern for step duration. Moreover, the number of errors in step duration is half that
of step length, regardless of treadmill belt speed. This is a strong indication that spatial and temporal controls of locomotion
are accessible through distinct neural circuits. This interpretation is corroborated by the earlier study of Malone and Bastian,
who investigated adaptation of spatial and temporal aspects of walking to a sustained perturbation, generated by a split-belt
treadmill.18 They demonstrated that conscious correction facilitates adaptation, whereas distraction slows it. The unexpected
finding of their study was that those manipulations affected the adaptation rate of the spatial elements of walking, but not of
the temporal ones. In the follow-up study Malone et al25 demonstrated that temporal and spatial controls of symmetric gait
can be adapted independently. Please note that continuous, conscious assessment of distance to surrounding objects lies at
the heart of the control problem of remaining stationary on a moving treadmill belt. Thus, the large number of errors in step
length as compared to step duration may reflect both the dominant role of spatial control and its susceptibility to distraction. It
is worth mentioning that in casual walking, the coefficient of variation of stride time is much smaller than that of stride length
and of walking speed.10
Step speed may be interpreted as the output of the intricate neuromuscular control system, which integrates different
4/11
sensory-motor processes. The ratio of average values of step length and frequency, or walk ratio, is constant over a broad
range of walking speeds. In other words, there is a linear relation between these gait parameters (the stride length -- cadence
relationship), a pre-programmed pattern which presumably simplifies gait control in steady state walking.26 Let us analyze the
interplay of step duration and step length during transient changes following the occurrence of errors. We previously pointed
out that these two parameters are controlled in distinct ways. In particular, the probability of evoking the interleg control of
step length is at least two times smaller than that of evoking the intraleg control (Table 1). In sharp contrast, the probability
of either inter- or intraleg control of step speeds is comparable. Thus, we may hypothesize that negative-feedback adjustment
of step duration of the contralateral leg underlies the interleg control of step speed. It is worth emphasizing that the intraleg
control of step speed is stronger than the interleg control.
The recent work of Dingwell et al.16 provides insight into the maintenance of speed during treadmill walking. A subject
can in principle choose any combination of stride length and time that yields step speed equal to that of a treadmill belt. These
pairs of values form in phase-space a diagonal line called a goal equivalent manifold (GEM).27 Dingwell et al. decomposed
deviation from this manifold into tangent and transverse components. Only the latter component was tightly controlled.
Moreover, the time series of transverse deviations exhibited statistical antipersistence characteristic of stride speed. This study
underscores the significance of interleg control of gait parameters. We believe that the GEM decomposition should be applied
to time series of step velocities to quantify the interleg control in a more sophisticated way.
In able-bodied gait, asymmetry in spatio-temporal and kinematic parameters (such as speed profiles, step and stride length,
foot placement angle, maximum knee flexion) for the left and right leg has been frequently reported.28 To the best of our
knowledge, the present study is the first observation of asymmetry in dynamics of human gait parameters. With the exception
of step duration control at the lowest speed, for all three gait parameters D DBE < 0, indicating dominance of the right leg in
short-term control. The origin of this asymmetry can be traced back to differences in the relative contribution of lower limbs
to control and propulsion -- the effect known as functional gait asymmetry.19 More specifically, the leg with greater muscle
power generation dominates propulsion, while the support and control functions are more conspicuous for the leg with greater
power absorption. Humans are typically right-footed for mobilization and left-footed for postural stabilization.
Special consideration should be given to step duration and step length control at the lowest speed v = 1.1m/s. Only in this
case, the asymmetry parameter D DBE was greater than zero, indicating the dominance of left lower limb. Note that the lowest
asymmetry, D DBE, was observed for all three parameters at v = 1.1m/s. Differences in low-speed gait have been reported
before. Terrier and Schutz29 demonstrated that during overground walking, at low speeds the majority of subjects adopted a
higher walk ratio and had a higher variability of stride time. However, in this study the lowest treadmill speed coincides with
the preferred walking speed (PWS) of young subjects.23 There are two possible explanations for the positive value of D DBE.
It is likely that in the vicinity of PWS priority is given to balance maintenance and consequently stride duration control is
shifted to the left leg, which is used for postural stabilization. Please note that our cohort included only subjects with clearly
pronounced right foot preference. Alternatively, reversed asymmetry for step duration and low values of D DBE for step
length and step speed may indicate that there exists a different strategy for control of gait parameters in overground walking
(treadmill walking at v = 1.1m/s may not be challenging for young subjects and may resemble unconstrained overground
walking). This argument is plausible because in motor coordination tasks, humans correct only those deviations that interfere
with task goals and allow variability in redundant (task-irrelevant) dimensions.31 Following the logic of this minimum inter-
vention principle, in treadmill walking, step speed must be tightly regulated. However, in overground walking, higher priority
may be given, for example, to balance control, which would affect the value of the asymmetry parameter D DBE. These two
qualitatively different strategies may also reflect other fundamental differences between overground and treadmill walking.
The rate at which the environment flows past the eyes seems to be an important mechanism for regulating walking speed.32, 33
More specifically, vision is used correctively to maintain walking speed at a value that is perceived to be optimal. For treadmill
walking, a discrepancy between observed and expected visual flow leads to a significant reduction (about 20%) of PWS,34
as well as the speeds of walk-run and run-walk transitions.32 It is worth pointing out that as far as kinetic and kinematic
parameters are concerned, treadmill and outdoor gaits are similar.35
The discovery of dependence of functional asymmetry in short-term control of gait spatio-temporal parameters on treadmill
speed was an unexpected outcome of this research. The elucidation of the transition from left-leg to right-leg dominance in
short-term control entails determination of the PWS for each subject. Further research is also needed to understand why the
probability of compensatory response for interleg control is close to 1 and is almost two times greater than that of intraleg
control (Table 3). Undoubtedly, such a strong difference indicates different roles these two mechanism play in control of
gait during treadmill walking. One may hypothesize that the primary goal of interleg control is maintenance of balance via
negative feedback from either leg while achieving specific goals such as matching the speed of the treadmill belt requires
intraleg adjustments.
During human locomotion, the legs act as two coupled oscillators.36 However, most studies disregard bilateral coordination
and synchronization dynamics37 -- 39 and focus on single-leg variability (stride time, length, speed). Herein we demonstrated
5/11
asymmetric short-term intra- and interleg control of spatio-temporal gait parameters. We believe that a better understanding
of these effects will not only pave the way for more realistic models of gait variability and control, but also help to refine
procedures used in rehabilitation of gait impairments.
Methods
We recruited 20 healthy students (10M/10F, mean(SD): age 22 yr (2), height 1.73 m (0.1), weight 71 (15) kg, BMI 23 (4)) of
the Wroclaw University of Science and Technology, who all signed an informed consent. The study was performed according
to the Declaration of Helsinki and the protocol was approved by the Ethics Committee of Wroclaw Medical University. The
subjects were screened to exclude those with a history of orthopedic problems, recent lower extremity injuries, any visible
gait anomalies, or who were taking medications that might have influenced their gait. We only enrolled subjects who used the
right leg to: kick a tennis ball, manipulate a tennis ball around a circle, make a first step, make a step after being pushed from
behind. These purely bilateral tasks are frequently incorporated into foot-preference inventories.40, 41 The protocol began with
a 5 min familiarization period of walking on a level motor-driven treadmill. Then each subject was asked to walk 400 m three
times at 1.1 m/s, 1.4 m/s i 1.7 m/s (4 km/h, 5 km/h and 6 km/h). The objective was to investigate control of gait parameters at
treadmill speeds equal to or greater than the PWS of young subjects. Therefore, the lowest speed was equal to the preferred
walking speed reported by Terrier and Deriaz23 and slightly smaller than the values determined by Dal et al.34 (1.19 m/s) and
Dingwell11 (1.22 m/s).
The gait parameters were extracted from the trajectories of the 30 mm optical markers attached to both shoes below the
ankle. The movements of those markers were recorded using an in-house motion capture system with a frame rate of 240
Hz and 720p resolution. The optical tracking was implemented in C++ (Visual Studio 2013) using OpenCV library. A heel
strike was defined as the point where the marker of the forward foot was at its most forward point during each gait cycle. A
step length was the distance between the ipsilateral and contralateral heel strikes. A step duration was equal to the elapsed
time between the ipsilateral and contralateral heel strikes. A step speed was calculated as the quotient of step length and step
duration. The group averaged number of steps taken per trial was equal to 593 (23) at 1.1 m/s, 508 (59) at 1.4m/s, and 456
(56) at 1.7 m/s.
In Fig. 1 we present a time series of step duration for treadmill walking at 1.1 m/s. The circle in this figure indicates step
duration that was longer than the mean value (represented in this figure by horizontal, thick, dotted line) by more than 3/2 of
standard deviation (the upper, horizontal, thin dotted gridline represents this threshold). It is apparent that the duration of the
step, which immediately follows the "error", suddenly decreases (this shorter interval is marked by the filled rectangle). This
example hints at the existence of an interleg control mechanism that stabilizes the stride interval.
0.56
0.54
0.52
0.5
0.48
]
s
[
n
o
i
t
a
r
u
d
p
e
t
s
compensation
0.46
60
error
left
right
65
70
75
80
85
90
time [s]
Figure 1. Time series of step durations for treadmill walking at 4 km/h. The circle indicates the step duration of left leg
which was longer than the mean value (represented in this figure by horizontal, thick, dotted line) by more than 3 ⁄ 2 of
standard deviation (the upper, horizontal, thin dotted gridline represents this threshold). It is apparent that the duration of the
step which immediately follows the "error" suddenly decreases.
Let N be the number of steps taken by each leg. Let us introduce a notation that facilitates the analysis of interleg control.
n=1 in the following
We write the time series of length 2N of one of the gait parameters (step duration, length or speed) {Sn}2N
form:
{Sn} = {I1,C1,I2,C2...IN,CN} ,
(1)
6/11
s I and s C are
where subseries(cid:8)I j(cid:9)N
standard deviations of these series. The simple moving averages (the unweighted mean of the previous m data) of (cid:8)I j(cid:9) and
(cid:8)Cj(cid:9) are denoted by ¯I(m) and ¯C(m), respectively.
j=1 correspond to the ipsilateral and the contralateral leg, respectively.
j=1 and (cid:8)Cj(cid:9)N
We define as errors these values Ii which satisfy all of the following criteria:
Ii − ¯I(m) > 1.5s I
Ii − Ii−1
(cid:12)
> 3%
(cid:12)
(cid:12)(cid:12)
(cid:12) < 0.5s C
Ci−1 − ¯C(m)(cid:12)
(cid:12)
(cid:12)
100%(cid:12)
(cid:12)
(cid:12)(cid:12)
Ii−1
(2)
(3)
(4)
These undoubtedly heuristic criteria are used to detect abrupt changes (equation (3)) which lead to conspicuous deviations
from the moving average value (equation (2)) and which are not brought about by a deviation in the preceding step of the
contralateral leg (equation (4)). As previously mentioned, we dub such events errors, but bear in mind that they may originate
either from the motor control system failure, or from the necessary adjustment of the subject's position on a treadmill. The
rationale for using the moving average in the above definition of an error stems from non-stationarity of gait time series. This
modification ensures that during transient linear trends the large deviation from the global mean value does not invoke the
detection algorithm. Please note that equation (3) by itself is another safeguard for false error detection caused by the transient
drift of local mean value. Herein, we report the values for m = 10.
to denote a deviation of a given gait parameter from its moving average value, e.g. D Ii = Ii − ¯I(m). In interleg
control the gait parameter of contralateral leg Ci changes in such a way as to decrease deviation of Ii + Ci. To quantify such
stabilization, we introduce the following metric:
Let us use D
Dinter
i =
D Ii + D Ci
D Ii + D Ci−1
.
(5)
The stabilization occurs when Dinter
two cases. In the first case the D Ci has the opposite sign to D Ii:
i < 1. The numerator in the above equation may become smaller than the denominator in
i = D IiD Ci < 0,
Sinter
(6)
in other words, the contralateral leg compensates for errors, as shown in Fig. 2a. The perfect compensation corresponds
to Dinter
= 0. In the alternative scenario, only the magnitude of the deviation of the contralateral leg from the mean value
decreases ( Sinter
i > 0) as illustrated by Fig. 2b.
i
Figure 2. In interleg control the gait parameter of contralateral leg Ci changes in such a way as to decrease deviation of
Ii +Ci from the mean value. a) The error of ipsilateral leg Ii is immediately compensated for by the contralateral leg. b) In the
alternative scenario, only the magnitude of the deviation of the contralateral leg from the mean value decreases.
It is possible that an error does not bring about a sudden change of gait parameter of contralateral leg.
In this case,
>= 1. However, stabilization may occur during the next step of ipsilateral leg. The change of Ii+1 may reduce the
Dinter
deviation D Ii+1 + D Ci. We refer to such a scenario as an intraleg control and define a corresponding metric:
i
Dintra
i =
D Ii+1 + D Ci
D Ii + D Ci
.
(7)
7/11
To be able to directly compare the properties of both types of control (inter- and intraleg) we distinguish whether the intraleg
control was achieved via compensation:
i = D IiD Ii+1 < 0,
Sintra
(8)
as shown in Fig. 3a, or by the reduction of the magnitude of the displacement of gait parameter of the ispislateral leg from the
moving average value (Fig. 3b).
Figure 3. The error in a gait spatio-temporal parameter may not evoke a sudden change of the parameter of contralateral
leg. However, the deviation of the next ipsilateral gait parameter Ii+1 from the mean value ¯I may decrease as a result of: a)
compensation or b) the reduction of the magnitude of the displacement of parameter of the ispislateral leg from the mean
value.
The flowchart in Fig. 4 elucidates the analysis of the dynamics of gait parameter time series which follows the occurrence
of errors. Using Dinter and Dintra, we detect the activation of inter- and intraleg control mechanism, respectively.
Figure 4. Flowchart of analysis of dynamics of gait parameter which follows an error -- gait-parameter's value which satisfy
the criteria described by equations (2-4).
In our analysis of the experimental data, we use a more specific notation for the interleg parameter Dinter. For example,
to indicate that an error in a given gait parameter of the left leg was followed by an adjustment of this parameter by the right
8/11
leg, we write DLR. In the same vein, we use DLL , DRR to denote intraleg leg control parameter for the left and right leg,
respectively.
In most cases D values, for a given gait parameter, speed, and control type, were not normally distributed (the Shapiro-
Wilk test). For a given speed and gait parameter, the Levene's test showed equality of variances among the control types (with
the exception of step duration at 1.1 m/s and step length at all speeds). For a given gait parameter and control type (L-L,
R-R, L-R, R-L), we investigated the dependence of D on treadmill speed. In this case, the Levene's test showed homogeneity
of variance. Consequently, the Kruskal-Wallis test with Tukey's post hoc comparisons was used to detect differences across
speed and control type. The significance threshold was set to 0.05.
To quantify functional asymmetry in control of gait spatio-temporal parameters we need to take into account the stochastic
aspect of motor control system. Let us employ an analogy of detailed balance equation of statistical physics42 and call it gait
detailed balance equation (DBE). In its original formulation, detailed balancing relates the relative population of two states by
the probability of a transition between them. The principle applies equally well to physical systems, mathematical probability
densities, or statistical processes in a variety of forms.
The smaller D the better stabilization of stride gait parameters. Consequently, the influence of a control mechanism (inter
or intra) on gait parameters is proportional to its probability of occurrence and the inverse of the corresponding mean value of
control parameter ¯D. For example, for the right lower limb, we may write:
DBER =
pRR
¯DRR +
pLR
¯DLR ,
in the same vein, for the left lower limb:
DBEL =
pLL
¯DLL +
pRL
¯DRL .
The perfect symmetry corresponds to the following equality:
DBEL = DBER.
We quantify the asymmetry in control of gait spatio-temporal parameters with the relative difference expression:
D DBE = 2
DBEL − DBER
DBEL + DBER
.
(9)
(10)
(11)
(12)
References
1. West, B. J. & Griffin, L. A. Biodynamics: Why the Wirewalker Doesn't Fall (Wiley, 2004).
2. Hausdorff, J. M., Peng, C. K., Ladin, Z., Wei, J. Y. & Goldberger, A. L. Is walking a random walk? Evidence for long-
range correlations in stride interval of human gait. Journal of applied physiology (Bethesda, Md. : 1985) 78, 349 -- 58
(1995).
3. Hausdorff, J. M. et al. Fractal dynamics of human gait: stability of long-range correlations in stride interval fluctuations.
Journal of applied physiology (Bethesda, Md. : 1985) 80, 1448 -- 57 (1996).
4. Bassingthwaighte, J. B., West, B. J. & Liebovitch, L. S. Fractal Physiology (Oxford University Press, 1994).
5. Hausdorff, J. M. Gait dynamics, fractals and falls: finding meaning in the stride-to-stride fluctuations of human walking.
Human movement science 26, 555 -- 89 (2007).
6. Zarrugh, M. Y., Todd, F. N. & Ralston, H. J. Optimization of energy expenditure during level walking. European journal
of applied physiology and occupational physiology 33, 293 -- 306 (1974).
7. Srinivasan, M. & Ruina, A. Computer optimization of a minimal biped model discovers walking and running. Nature
439, 72 -- 75 (2006).
8. Srinivasan, M. & Ruina, A. Idealized walking and running gaits minimize work. Proceedings of the Royal Society A:
Mathematical, Physical and Engineering Sciences 463, 2429 -- 2446 (2007).
9. Diedrich, F. J. & Warren, W. H. Why change gaits? Dynamics of the walk-run transition. Journal of experimental
psychology. Human perception and performance 21, 183 -- 202 (1995).
10. Kito, T. & Yoneda, T. Dominance of gait cycle duration in casual walking. Human movement science 25, 383 -- 92 (2006).
9/11
11. Dingwell, J. B. & Marin, L. C. Kinematic variability and local dynamic stability of upper body motions when walking at
different speeds. Journal of biomechanics 39, 444 -- 52 (2006).
12. Zehr, E. P. & Stein, R. B. What functions do reflexes serve during human locomotion? Progress in neurobiology 58,
185 -- 205 (1999).
13. Warren, W. H., Kay, B. A., Zosh, W. D., Duchon, A. P. & Sahuc, S. Optic flow is used to control human walking. Nature
neuroscience 4, 213 -- 6 (2001).
14. Bent, L. R. et al. When is vestibular information important during walking? Journal of neurophysiology 92, 1269 -- 1275
(2004).
15. Rossignol, S., Dubuc, R., Gossard, J.-P. & Dubuc, J. Dynamic Sensorimotor Interactions in Locomotion. Physiological
Reviews 86, 89 -- 154 (2006).
16. Dingwell, J. B., John, J. & Cusumano, J. P. Do humans optimally exploit redundancy to control step variability in walking?
PLoS computational biology 6, e1000856 (2010).
17. Reisman, D. S., Block, H. J. & Bastian, A. J. Interlimb coordination during locomotion: what can be adapted and stored?
Journal of neurophysiology 94, 2403 -- 2415 (2005).
18. Malone, L. a. & Bastian, A. J. Thinking about walking: effects of conscious correction versus distraction on locomotor
adaptation. Journal of neurophysiology 103, 1954 -- 1962 (2010).
19. Sadeghi, H., Allard, P. & Duhaime, M. Functional gait asymmetry in able-bodied subjects. Human Movement Science
16, 243 -- 258 (1997).
20. Terrier, P., Turner, V. & Schutz, Y. GPS analysis of human locomotion: further evidence for long-range correlations in
stride-to-stride fluctuations of gait parameters. Human movement science 24, 97 -- 115 (2005).
21. Deligni`eres, D. & Torre, K. Fractal dynamics of human gait: a reassessment of the 1996 data of Hausdorff et al. Journal
of applied physiology (Bethesda, Md. : 1985) 106, 1272 -- 1279 (2009).
22. West, B. J. & Scafetta, N. Nonlinear dynamical model of human gait. Phys. Rev. E 67, 051917 (2003).
23. Terrier, P. & D´eriaz, O. Persistent and anti-persistent pattern in stride-to-stride variability of treadmill walking: influence
of rhythmic auditory cueing. Human movement science 31, 1585 -- 97 (2012).
24. Snaterse, M., Ton, R., Kuo, A. D. & Donelan, J. M. Distinct fast and slow processes contribute to the selection of preferred
step frequency during human walking. Journal of applied physiology (Bethesda, Md. : 1985) 110, 1682 -- 1690 (2011).
25. Malone, L. A., Bastian, A. J. & Torres-Oviedo, G. How does the motor system correct for errors in time and space during
locomotor adaptation? Journal of Neurophysiology 108, 672 -- 683 (2012).
26. Egerton, T., Danoudis, M., Huxham, F. & Iansek, R. Central gait control mechanisms and the stride length - cadence
relationship. Gait and Posture 34, 178 -- 182 (2011).
27. Cusumano, J. P. & Cesari, P. Body-goal variability mapping in an aiming task. Biological Cybernetics 94, 367 -- 379
(2006).
28. Sadeghi, H., Allard, P., Prince, F. & Labelle, H. Symmetry and limb dominance in able-bodied gait: a review. Gait &
posture 12, 34 -- 45 (2000).
29. Terrier, P. & Schutz, Y. Variability of gait patterns during unconstrained walking assessed by satellite positioning (GPS).
European Journal of Applied Physiology 90, 554 -- 561 (2003).
30. Terrier, P. Fractal Fluctuations in Human Walking: Comparison Between Auditory and Visually Guided Stepping. Annals
of biomedical engineering 44, 2785 -- 2793 (2016).
31. Todorov, E. & Jordan, M. Optimal feedback control as a theory of motor coordination. Nature neuroscience 5, 1226 -- 1235
(2002).
32. Mohler, B. J., Thompson, W. B., Creem-Regehr, S. H., Pick, H. L. & Warren, W. H. Visual flow influences gait transition
speed and preferred walking speed. Experimental Brain Research 181, 221 -- 228 (2007).
33. O'Connor, S. M. & Donelan, J. M. Fast visual prediction and slow optimization of preferred walking speed. Journal of
Neurophysiology 107, 2549 -- 2559 (2012).
34. Dal, U., Erdogan, T., Resitoglu, B. & Beydagi, H. Determination of preferred walking speed on treadmill may lead to
high oxygen cost on treadmill walking. Gait & posture 31, 366 -- 9 (2010).
10/11
35. Riley, P. O. et al. A kinematics and kinetic comparison of overground and treadmill running. Medicine and science in
sports and exercise 40, 1093 -- 100 (2008).
36. Beek, P. J., Peper, C. E. & Stegeman, D. F. Dynamical models of movement coordination. Human Movement Science 14,
573 -- 608 (1995).
37. Bartsch, R. et al. Fluctuation and synchronization of gait intervals and gait force profiles distinguish stages of Parkinson's
disease. Physica A: Statistical Mechanics and its Applications 383, 455 -- 465 (2007).
38. Krasovsky, T. et al. Stability of gait and interlimb coordination in older adults. Journal of Neurophysiology 107, 2560 --
2569 (2012).
39. Wuehr, M., Pradhan, C., Brandt, T., Jahn, K. & Schniepp, R. Patterns of optimization in single- and inter-leg gait
dynamics. Gait & posture 39, 733 -- 8 (2014).
40. Chapman, J. P., Chapman, L. J. & Allen, J. J. The measurement of foot preference. Neuropsychologia 25, 579 -- 584
(1987).
41. Gabbard, C. & Hart, S. A question of foot dominance. The Journal of general psychology 123, 289 -- 296 (1996).
42. Kampen, N. V. Stochastic Processes in Physics and Chemistry, Third Edition (North Holland, 2007), 3 edition edn.
Author contributions statement
M.L and B.J.W conceived the experiment, K.K. conducted the experiment, all authors analyzed the results and contributed to
the manuscript.
Additional information
Competing financial interests There were no conflicts of interest.
11/11
|
1104.4201 | 1 | 1104 | 2011-04-21T08:31:48 | Self-organised criticality in base-pair breathing in DNA with a defect | [
"physics.bio-ph",
"cond-mat.soft",
"math.DS",
"nlin.AO",
"q-bio.BM"
] | We analyse base-pair breathing in a DNA sequence of 12 base-pairs with a defective base at its centre. We use both all-atom molecular dynamics (MD) simulations and a system of stochastic differential equations (SDE). In both cases, Fourier analysis of the trajectories reveals self-organised critical behaviour in the breathing of base-pairs. The Fourier Transforms (FT) of the interbase distances show power-law behaviour with gradients close to -1. The scale-invariant behaviour we have found provides evidence for the view that base-pair breathing corresponds to the nucleation stage of large-scale DNA opening (or 'melting') and that this process is a (second-order) phase transition. Although the random forces in our SDE system were introduced as white noise, FTs of the displacements exhibit pink noise, as do the displacements in the AMBER/MD simulations. | physics.bio-ph | physics |
Self-organised criticality in base-pair
breathing in DNA with a defect
Ciprian-Ionut¸ DUDUIAL A, Jonathan A.D. WATTIS
School of Mathematical Sciences, University of Nottingham,
Nottingham, United Kingdom, NG7 2RD, UK
Charles A. LAUGHTON,
School of Pharmacy, University of Nottingham,
Nottingham, United Kingdom, NG7 2RD, UK
November 4, 2018
Abstract
We analyse base-pair breathing in a DNA sequence of 12 base-pairs with a defective
base at its centre. We use both all-atom molecular dynamics (MD) simulations and
a system of stochastic differential equations (SDE). In both cases, Fourier analysis of
the trajectories reveals self-organised critical behaviour in the breathing of base-pairs.
The Fourier Transforms (FT) of the interbase distances show power-law behaviour with
gradients close to −1. The scale-invariant behaviour we have found provides evidence
for the view that base-pair breathing corresponds to the nucleation stage of large-scale
DNA opening (or 'melting') and that this process is a (second-order) phase transition.
Although the random forces in our SDE system were introduced as white noise, FTs
of the displacements exhibit pink noise, as do the displacements in the AMBER/MD
simulations.
Keywords: DNA, breathers, self-organised criticality, stochastic differential equations,
1/f noise.
PACS 05.10.Gg - Stochastic models in statistical physics and nonlinear dynamics
PACS 89.75.Fb - Self-organization in complex systems,
PACS 05.65.+b - Self-organised criticality
PACS 87.14.gk - DNA
MSC 60G18 Probability theory and stochastic processes: self-similar processes
MSC 37K40 Dynamical systems: Soliton theory, asymptotic behavior of solutions
MSC 92C45 Biology and other natural sciences: Kinetics in biochemical problems
1
Introduction
Much has been written about the process by which double-stranded DNA becomes two sepa-
rated single strands, see, for example, (1; 2). Several authors refer to the melting of DNA as
1
having the form of a phase transition, in which the opening of one or a few base-pairs is akin
to nucleation, and the subsequent growth of open 'bubbles' is similar to the growth of crystal
nucleii. This two-stage process is relatively well understood in the contexts of crystal growth,
and aerosol formation, but less so in kinetics of DNA replication. The aim of this paper is to
show that the nucleation event -- that is the initial opening of bases -- exhibits self-similar, or
scale-free, critical behaviour, as one would expect at a phase transition. Whilst other studies
have analysed bubble growth and the statistics of bubble length, we choose to focus on the
temporal statistics. Our model is more detailed, only focusing on 12 base-pairs and the open-
ing of the first base; our model is thus significantly smaller than the bubble growth models of
(1; 2); however, our models are more accurate in that the AMBER simulations (3) (Assisted
Model Building with Energy Refinement) include the effect of every atom in the DNA and
the water molecules in the environment, and our SDE models include an accurate fitting of
the nonlinear inter-base potential energy as described in our earlier work (4).
According to Watson & Crick (7), the structure of a DNA duplex consists of two chains
of bases. These bases are of four types: the purines Adenine (A) and Guanine (G) and the
pyrimidines Cytosine (C) and Thymine (T). Along the chains, the bases are linked by covalent
bonds, while the opposite bases from the two chains pair together by two or three hydrogen
bonds forming base-pairs. Only A-T or C-G pairs are possible. Given this information, we
use a lattice representation for our DNA sequence, as illustrated in Figure 1. Breathing -- the
localised separation of complementary bases -- takes place on the microsecond timescale in
normal DNA, which is beyond the range of all-atom molecular dynamics (MD) packages. The
insertion of a defect, that is, replacing a thymine (T) with a difluorotoluene (F) base at the
lattice site n = 0, increases the frequency of breathing due to the weakening of the inter-chain
potential. This makes breathing occur on the nanosecond timescale and hence it becomes
accessible to MD simulation techniques. Biologically the reason for considering the inclusion
of a defect is to study fidelity, and the effects of errors, in DNA replication. The defect F has
been considered previously, for example, by Cubero et al. (8), who considered such a system
without any externally imposed twist. It is possible that proteins may locally alter the twist
of a DNA helix in order to ease the process of localised melting. Hence, here, we impose a
twist on the DNA structure in order to investigate its effect on base-pair breathing.
s
s
s
s
s
s
u−2(t) u−1(t)
✇
❣
✇
γ
k
✟✟✟✟
u−3(t)
γ
✟✟✟✟
k
v−3(t)
❣
✇
❣
k
γ
✇
❅
k
❅
❣
v−2(t)
❅
✇
❅
❣
❅
❅
u0(t)
✡
❣
✡
❅✡
✇
E0
bk
bk
bk
bk
❣
❣
❅❍❍❍❍
✇
v−1(t)
v0(t) v1(t)
✇
✇
✇
✡✡❅
✡
u1(t)
❅
❣
k
γ
k
❣
❅
u2(t)
❅
❣
✇
γ
❣
✇
v2(t)
s
s
s
s
s
s
Figure 1: IIllustration of the mesoscopic model of DNA.
In this paper, we analyse base-pair breathing at a defect in double-stranded DNA parame-
terised using the molecular dynamics (MD) simulation tool AMBER. We consider a range
of helicoidal twist angles from 30◦ to 40◦ per base-pair at rest, including the typical twist
of 36◦. In (4) we derived a coarse-grained model based on stochastic differential equations
with one variable for the displacement of each base from its equilibrium position. This model
2
includes explicit random forcing terms, which model the interactions of water molecules with
the bases, and the whole model has been parameterised to the all-atom MD simulations de-
rived using AMBER. Developed from the models of (5; 6), the incorporation of noise and
damping terms enables us to investigate the temporal dynamics of breathing. The model is
fitted to AMBER data using a sophisticated maximum likelihood estimation procedure which
is described in (4) and summarised in Section 2.2 here. The associated fluctuation-dissipation
relation, the dependence of parameters on angle of twist and form of the inter-base potential
are also discussed in (4).
The DNA sequence under study consists of 12 base-pairs and we have taken that the DNA
sequence is surrounded by a water box. The analysis of the simulations of a DNA molecule
obtained using AMBER and our SDE system presented in (4) revealed that the amplitude of
fluctuations is slightly reduced in the SDE model, but the breathing length and frequency are
similar. In addition, the derivation of the SDE model requires us to modify the fluctuation-
dissipation relation in the reduced, or coarse-grained, mesoscopic models. We also make a
distinction between the potential of mean force (the free energy) and the various potential
energies in our system. This is supported by an analysis of the importance of the damping
term in preserving the system energy (which, in combination with the forcing noise terms
gives rise to the entropic component of the free energy) and the way in which the along-
chain interactions influence the length of a breathing event. We have also shown in (4) that
breathing events are due not only to inhomogeneities in the inter-strand interactions, but
also to a significant change in along-chain interactions and the helical twist of the DNA,
which potentially influences the interactions between the DNA molecule and the surrounding
solvent.
A more detailed analysis of our simulations presented below reveals an interesting result:
rather than a well-defined breathing frequency that depends on the twist angle via the energy
barrier between open and closed states of the central basepair, we find that at all angles of twist
the DNA exhibits breathing across a wide range of frequencies and the amplitude-frequency
relationship exhibits scale-free behaviour. Most previous results in the literature show that
the energy is transferred between nonlinear localized modes with particular frequencies and
between such modes and phonons, as discussed by Peyrard et al. (9; 10; 11) or as shown by
the results of Gaeta et al. (12; 13; 14), for example.
In the rest of this section we review some of the relevant theory and applications involv-
ing self-organised systems.
In Section 2 we outline the models we use and the simulation
techniques. Section 3 contains the results of both approaches, showing the consistency of
outcomes. Section 4 concludes the paper with a summary and discussion.
1.1 Self-organisation and criticality
Many dynamical systems evolve to a steady state (or an equilibrium) solution or to a limit
cycle. More complicated large-time behaviour includes spatio-temporal chaos which is often
characterised by strange attractors in phase space (15). Through the study of cellular au-
tomata, Wolfram (16) characterised large time behaviour into four classes. He empirically
identified the following qualitative classes: spatially homogeneous systems (akin to steady-
states), periodic structures (limit cycles), systems with chaotic aperiodic behaviour (chaos)
3
and, most notably, a fourth, that of complicated localised and possibly propagating structures.
We claim that this last category is the most appropriate classification for the behaviour which
we observe later in this paper, and, more precisely, this behaviour corresponds to self-organised
criticality (SOC).
In physics, a critical point specifies the conditions, such as temperature, pressure or composi-
tion, at which a phase boundary is not valid anymore. Here, by 'phase' we understand a state
of a system for which the physical properties of a component are uniform. As one approaches
the critical point, the properties of the different phases approach each other. In other words,
a critical point refers to a system configuration to which the system evolves without ever
approaching a fixed equilibrium state. Mathematically, one can define characteristic sizes
of behaviour (lengths or timescales of events) which describe the system. When these sizes
become infinite the system is classed as critical, that is, fluctuations occur at all length scales.
Systems having the SOC property present a spatially or temporally scale-invariant behaviour
without the need to tune any parameter to a specific value. This ubiquity of scale-free
behaviour (18) in such models shows that complex behaviour can be stable. This contrasts
with systems where one would generically expect some stable steady behaviour for a wide
range of parameter values, and as a parameter changes, one might observe either a smooth
change in the system's behaviour or a bifurcation to a qualitatively different steady-state or
limit cycle, for example. In the special case when the parameter takes on the threshold value
between two states, more complex phenomena may be observed. This is a typical form of
behaviour around a phase transition. In general, the total number of states is finite and the
transitions can be characterised using a cellular automaton structure (19).
When analysing a large system, we aim to reduce its complexity to a few degrees of freedom, for
which the coupling can be defined in a general manner and hence we obtain some averaged,
or coarse-grained, behaviour over ignored quantities and includes corresponding averaged
interactions within the system and the surrounding environment. For dynamical systems,
such a dimensional reduction can be achieved by the "slaving principle" (17) which leads
to the the study of low-dimensional attractors. This is often a straightforward method. For
example, "fast modes" at equilibrium can be slaved to a few slowly-evolving modes. However,
sometimes a system responds on both fast and slow timescales, even at large times, and we
require an alternative theory, such as the idea of self-organised systems, whose behaviour
cannot be explained using the slaving principle or other reductions.
Some attracting critical points of dynamical systems can be characterised using the concept
of self-organized criticality (SOC), which was first introduced by Bak et al. (22; 23). Using
simple automata, they demonstrated power-law relationships and 1/f noise (also known as
"flicker noise") in spatially extended systems, this behaviour illustrates critical phenomena,
and underpins more general scale-invariant behaviour and fractals. They studied the dynamics
of damped pendula and the slope of sandpiles, determining critical points of the systems.
One aspect of self-organised criticality is the separation of timescales:
in the most familiar
application of sandpiles, grains are continually added on a faster timescale. The gradient
of the pile slowly steepens and there are avalanches (large-scale reorganisation of the pile)
which take place rapidly but are separated by large time intervals (relative to the timescale
at which grains are added to the pile). Bak et al also noted that changing the values of
system parameters did not affect the emergence of critical behaviour. Avalanches in a one-
dimensional sandpile are also analysed by Chapman et al. (see, for example, (24)) who showed
4
that the distribution of energy discharges due to internal reorganizations have a power-law
form and so demonstrate that the system is self-organized.
In general, for a noisy system the power spectrum has the form S(f ) = cf −β, where c is a
constant. The noise present in the system can be classified in three important categories as
follows:
• white noise, for β = 0;
• pink noise, for β = 1;
• red noise (also known as Brownian noise), for β = 2.
However, the term "1/f noise" is widely used to refer to any noise with a power spectral
density S(f ) ∝ f −β, with 0 < β < 2. For 1/f noise that occurs in nature, β is usually close
to 1.
Although there is no single well-defined class of systems having the SOC property, it is typ-
ically observed in complex systems with slowly-driven nonequilibrium behaviour, for which
the causes of an event taking place in a system cannot be explained simply through some
parameter values. Several studies of SOC show that scale-invariant phenomena can be deter-
mined at critical points, but not necessarily at any critical point. There are two important
categories of such phenomena: fractals (20) and power laws (21). Whilst the first category
involves geometric shapes, which can be split into parts that are reduced-size copies of the
initial shape, the second deals with frequency-dependent quantities and, hence, is relevant in
the analysis of some Hamiltonian systems. We note that self-organised systems are always at
criticality, but not all critical systems are self-organised.
The range of systems exhibiting critical properties varies from earthquakes (25; 26) and forest-
fires (27; 28) to biological systems, such as proteins (29; 30), the brain (31) and even DNA.
Selvam (32), for example, studies the distribution of bases in a human DNA sequence and
shows that the C-G base-pair frequency distribution exhibits a universal inverse power-law
form. Also, Harris et al. (33) analyse the configurational entropy of a DNA molecule based
on the entropy estimation for a Gaussian configuration given by Schlitter (34), which helps
investigate whether a steady state has been reached during a simulation. They show that
the estimate of the entropy Sn depends on the number of data points n and this relation is a
power law (with exponent between zero and minus one).
2 Modelling DNA
In our system, we also have a separation of timescales: there are rapidly oscillating forcing
terms applied to the DNA chain illustrated in Figure 1 (these model the interactions with
water molecules and are akin to grains being added to the sandpile) and the occasional larger-
scale restructuring of the chain as the base-pairs open or close at the start and end of breathing
events (which are akin to avalanches in sandpiles). We now discuss in detail the modelling
approaches we adopt.
5
2.1 All-atom MD modelling using AMBER
As already mentioned, we obtain data from an all atom simulation of DNA using the package
AMBER (3). The DNA sequence analysed contains 12 base-pairs as follows:
C T T T T G F A T C T T
G A A A A C A T A G A A
This sequence is analysed at a constant temperature of T = 293K, in the presence of a
surrounding water box. The solvent has to be taken into account because it influences the
displacement of atoms and through other bonds, affects the hydrogen bonds linking the bases
on the DNA strands (36). Even if the breathing events occur on the nanosecond time-scale and
the DNA sequence contains only 12 base-pairs, which together with the sugars and phosphate
groups represent 763 atoms, the number of degrees of freedom in our system is actually
very large (16682) due to the water box. This means most of the time is spent computing
information about the solvent, even though this information is not used for our analysis, since
we focus only on the DNA bases and their dynamics.
Note that AMBER considers that the normal twist by default is about 32.5◦. In order to
avoid this inconvenience, we have constructed the DNA sequence by considering the degree of
twist at rest. The degree of twist is preserved by imposing a harmonic restraint on the atoms
at the end bases. More precisely, we have considered a constant energy (of 1 kcal mol−1 A−2)
and hence a constant force acting on the end bases, in order to keep the DNA atoms close to
their initial positions. Applying this restraint to the end bases allows the A-F pair to breathe
and so explore a larger volume of phase space than the other base-pairs.
The force field used during the AMBER simulations is also important. For a DNA molecule,
the predefined FF99SB all-atom force field was used. Moreover, AMBER provides several
water models for residues with name WAT -- the default is TIP3P, which was used in our
simulations. After creating the topology and coordinates files using the AMBER packages
LEaP and nucgen, we have used SANDER for energy minimization. This process involves
a structural relaxation, which is necessary because the coordinates file contains some initial
values that do not guarantee a minimum energy configuration, this reduces the possibility
of having conflicts or overlapping atoms. In addition to energy minimization, we have also
performed a few equilibration and MD simulations in which temperature changes.
Finally, our system is simulated using SANDER. Next, ptraj is used to measure the distance
between the A-F base-pair as well as the separations of the other base-pairs and the corre-
sponding velocities. This data is output every 1 ps or 2 fs depending on which simulation
study is being performed.
2.2 Stochastic mesoscopic model
In DNA, the rapid external fluctuation events correspond to random collisions of the bases
with the water molecules surrounding the biopolymer. These events correspond to pertur-
bations to the variables yn(t), fluctuations in these displacements may propagate along the
6
70
60
50
40
30
20
10
y
g
r
e
n
e
l
a
i
t
n
e
t
o
p
PMF(y0)
U(y0)
E0(y0)
0
−1
0
1
2
4
5
6
7
3
y0
Figure 2: The interchain interaction potential E0(y0), the total potential energy U(y0) =
0, and the potential of mean force P MF (y0) = U(y0) + √kBT η0y0 used in (4),
E0(y0) + 1
for a 36◦ twisted DNA sequence.
2bky2
chain, perturbing neighbouring bases, eventually causing the central base y0(t) to cross a
barrier in the potential E0(y0), see Figure 2.
A second source of data is the reduced model introduced in (4) which is based on a system
of stochastic ordinary differential equations (SDEs). We start from the deterministic system
H = X
n + 1
2 v2
n + 1
2γ(un − vn)2 + 1
(cid:8) 1
2 u2
n
2(bk − k)(cid:2)(u1 − u0)2 + (u0 − u−1)2 + (v1 − v0)2 + (v0 − v−1)2(cid:3)
+ 1
+E0(y0) − 1
2k(un+1 − un)2 + 1
2γ(u0 − v0)2,
2 k(vn+1 − vn)2(cid:9)
(1)
which is illustrated in Figure 1. Here, un(t) represent the deviations from equilibrium of the
bases on one chain of the DNA, and vn(t) the displacements of the bases on the second chain,
with the index n determining the position down each chain (−6 ≤ n ≤ 5). Since we are
interested in investigating in more detail a system similar to that of Guckian et al. (37) we
place a defect at the centre of the chain, that is, the u0 and v0 bases correspond to the A-F
base-pair. The parameter k describes the stiffness of the backbone down each side of the
double-helix structure, whist the parameter γ indicates the strength of interaction between a
base on one strand and its complement. These forces are assumed to be uniform along the
DNA double helix, except at the defect where they are replaced by the parameter bk and the
nonlinear force E ′
0(y0) respectively. To these we add noise and damping terms with coefficients
ǫ∗ and η∗ respectively. The dependence of all these parameters on the twist angle θ has been
determined in an earlier paper (4). We make the transformation yn = un − vn so as to obtain
7
equations of motion for the distances between base-pairs
dyn
dt
+ ǫξn,
d2yn
dt2 = k(yn+1−2yn+yn−1) − γyn − η
d2y−1
dt2 = bk(y0−y−1) − k(y−1−y−2) − γy−1 − η
d2y0
dt2 = bk(y1−2y0+y−1) −
(y0) − η0
d2y1
dt2 = k(y2−y1) − bk(y1−y0) − γy1 − η
dy0
dt
dy1
dt
dE0
dy
dy−1
dt
+ ǫ0ξ0,
+ ǫξ1.
(n > 1),
+ ǫξ−1,
(2)
(3)
(4)
(5)
The functions ξn(t) are white noise forcing terms. The quantities γ, k, bk, ǫ, ǫ0, η, η0 are
all fitted using a maximum likelihood estimation procedure, as is the interaction potential
E0(y0), a full description of this is given in (4; 35). A typical example of the multiwelled
interaction potential, E0(y0) is given in Figure 2. Note that the total potential energy of the
central defective base-pair is U(y0) = E0(y0) + 1
0, and the potential of mean force (free
energy) is given by P MF (y0) = U(y0) + √kBT η0y0, can easily be obtained from simulations
by plotting a histogram of binned displacement data.
2bky2
3 Results
We measure the distance between the two bases (A, F) at the defect over many nanoseconds,
that is, we sample y0(t) at specific intervals. Using data from both AMBER and the SDE
system we analyse the (discrete) Fourier transforms of y0(t) for a range of twist angles from
30◦ to 40◦ per base-pair.
The discrete Fourier transform is taken of the data, in the figures displayed later, we use the
notation DF T (ω) for by0(ω), and the hypothesis we are testing is that over a wide range of ω,
(6)
logby0(ω) = −β log ω + log C.
The highest frequency attainable is ωmax = 2π/2∆t, whilst the lowest frequency is given by
2π/T where T is the length of the simulation and ∆t is the sampling interval. Our initial
results are simulations of length approximately T = 10 ns sampled every ∆t = 1 ps (giving
∼ 104 data points). We then perform simulations of T = 2 ns sampled on a much finer scale
of ∆t = 2 fs (giving ∼ 106 data points). These are performed using both AMBER and our
SDE system. The SDE system is then subjected to a longer-time simulation of T = 100 ns,
sampled every ∆t = 1 ps (∼ 105 data points). This enables a wide range of frequencies to be
sampled: for the initial simulations 0.00063 < ω < 3.14 ps−1 (−7.4 < log ω < 1.14), for the
more frequently sampled simulations 0.0032 < ω < 1600 (−5.7 < log ω < 7.4), and for the
long simulations 2π × 10−5 < ω < 3.14 (−9.7 < log ω < 1.14).
Other frequencies which might be of relevance in interpreting the results are the upper and
lower limits of the phonon band (which we refer to as ωopt and ωac respectively), and the
frequency of the defect mode. Assuming that the defect mode can be approximated by
0 (0) = 14 and ωdef = 3.7, hence log ωdef = 1.3. Since 120 < γ < 165
y0 = −E ′′
and 160 < γ + 4k < 210, we have ωac = 11 and ωopt = 14.5, implying log ωac = 2.4 and
log ωopt = 2.7. These all lie well above the range of frequencies that we shall be interested in
below.
0 (0)y0 we find E ′′
8
3.1
Initial results
)
)
ω
(
T
F
D
(
g
o
l
(a)
10
5
0
−8
−6
0
2
−4
−2
log(ω)
(b)
10
5
)
)
ω
(
T
F
D
(
g
o
l
0
−8
−6
−4
−2
log(ω)
0
2
Figure 3: Discrete Fourier transforms (power spectra) of y0(t) plotted against ω on a log-log
scale. Data sampled every 1 ps for 10 ns from a simulation of a 38◦ twisted DNA sequence
from (a) AMBER, and (b) the SDE system (2) -- (5).
In Figure 3, we present the log-log plot of the discrete Fourier transform (DFT) against the
frequency ω, for 213 data points, that is, y0(t) sampled every 1 ps for 8 ns, for a 38◦ overtwisted
DNA sequence. This Figure shows a straight line fit over several orders of magnitude of ω
(−7 < loge ω < −1), with gradients of 0.7 -- 0.9 which are close to −1. Similar results are
obtained for a 34◦ twisted DNA sequence, for example, as can be seen in Figure 4. These
results suggests that there are breathing events of, and separated by, arbitrarily large times.
Thus, if a DNA strand was successively observed for increasingly long intervals of time, there
would always be breathing events of duration comparable to the total observation time. The
gradients of the logeDFT(ω) against loge ω lines for the full range of twist angles tested are
summarised in Table 1 and suggest the presence of generalised 1/f noise in our data (that is,
1/f β with 0 < β < 2).
Initially our aim was to identify the dominant frequencies of breathers at the defect through
the interchain distances. The asymptotic results of (6) initially appear to suggest that
breathers are time-periodic modes with well-defined frequencies; however, that theory actually
predicts a one-parameter family of "in-phase" breather modes with frequencies occupying the
9
)
)
ω
(
T
F
D
(
g
o
l
)
)
ω
(
T
F
D
(
g
o
l
(a)
−6
−4
−2
log(ω)
(b)
0
2
10
5
0
−8
10
5
0
−8
−6
−4
−2
log(ω)
0
2
Figure 4: Discrete Fourier transforms (power spectra) of y0(t) plotted against ω on a log-log
scale. Data sampled every 1 ps for 10 ns from a simulation of a 34◦ twisted DNA sequence
from (a) AMBER, and (b) the SDE system (2) -- (5).
Angle βAM BER
30◦
32◦
33◦
34◦
35◦
36◦
38◦
40◦
0.725
0.700
0.725
0.750
0.750
0.775
0.700
0.700
βSDE
0.750
0.725
0.775
0.825
0.775
0.825
0.875
0.700
Table 1: The gradient β of the log-log representation of DF T (y0) against ω, from ∼ 10 ns of
data, sampled every 1 ps obtained from the AMBER and SDE models.
full range of values from the bottom of the phonon band down to arbitrarily small frequencies
(as well as a one-parameter family of "out of phase" breathers with frequencies above the
top of the phonon band). Since the part of the frequency range that we are interested in
here is the small-ω limit, it is the former, in-phase, family that concerns us here. We assume
10
that a combination of the stochastic forcing noise and nonlinear interactions of phonons with
the breathers which changes their frequency over time and may even sporadically create and
destroy the breather modes.
3.2 More refined results
By decreasing the sampling interval (∆t), we expect to obtain more accurate results; the cost
being the increase in data storage requirements. Analysing data sampled every 2 fs for 2.1
ns (more precisely, 220 or 106 data points), we obtain qualitatively similar results, as can be
seen in the results presented in Figure 5, which shows the Fourier power spectra for a DNA
helix twisted to 34◦.
(a)
)
)
ω
(
T
F
D
(
g
o
l
)
)
ω
(
T
F
D
(
g
o
l
15
10
5
0
−5
−6
15
10
5
0
−5
−6
4
6
8
−4
−2
0
2
log(ω)
(b)
−4
−2
0
2
log(ω)
4
6
8
Figure 5: Log-log plots of the discrete Fourier transforms (power spectrum) from more refined
samplings of y0(t), specifically, data was sampled every 2 fs for 2.1 ns. Results for a 34◦
undertwisted DNA sequence from (a) AMBER and (b) the SDE system (2) -- (5).
As with the results shown in Section 3.1, the full range of twist angles from 30◦ to 40◦ have been
simulated and analysed the results are summarised in Table 2. These results, with the reduced
sampling interval, suggest that the average value of β for such AMBER simulations increases
to 0.93, which is much closer to unity than the values obtained when the sampling interval
is ∆t = 1ps (Table 1). Analysing a similar data set from the SDE model with more frequent
11
sampling shows a similar increase in β, from an average of 0.79 when ∆t = 1ps (Table 1)
to ∆t = 0.91 (Table 2). These results again confirm that our reduced mesoscopic model
accurately reproduces the self-organised DNA behaviour observed in the all-atom AMBER
simulation. This observed increase in β suggests that we might reasonably expect β → 1 as
∆t → 0 in both the AMBER and the SDE simulations.
Angle βAM BER
30◦
32◦
33◦
34◦
35◦
36◦
38◦
40◦
0.920
0.920
0.900
0.940
0.930
0.930
0.940
0.950
βSDE
0.900
0.895
0.910
0.920
0.900
0.910
0.940
0.905
Table 2: The β exponents derived from more refined sampling of y0(t), specifically every 2 fs
for 2.1 ns. The β-exponent is the gradient of the log-log representation of DF T (y0) against
ω, from AMBER and SDE simulations (compare with Table 1).
At higher values of ω, both AMBER and SDE simulations show changes in behaviour. Firstly,
in both cases, the line broadens as more points are plotted at higher values of log ω and
these display a greater variation; see the ranges −1 < log ω < 2 in Figures 3 and 4 and
−1 < log ω < 6 in Figure 5. Note that the same ranges apply to both AMBER simulations
and SDE simulations. Secondly, at larger ω, there is a more significant reduction in the
discrete Fourier transform, for the AMBER simulation this occurs from ω = 2 upwards,
following by a more abrupt decrease around ω = 6.5 in Figure 5(a); whereas, in the SDE
simulation, Figure 5(b), the spectrum has a simpler form with a more rapid linear decrease
with a larger gradient beyond ω = 2.5.
3.3 Long-time results
We have analysed a longer simulation of 100 ns, sampling y0(t) every ∆t = 1 ps using just
the SDE system; such a long simulation is beyond the scope of AMBER on currently avail-
able computing facilities. This length of simulation allows lower breathing frequencies to be
sampled, as shown in Figures 6 and 7. Here we observe the same self-organised behaviour as
in earlier graphs, although with some deviation from the straight line at particularly small
frequencies, namely those in the range −9 < log ω < −7.
For example, from the 216(∼ 105) data points in the simulation of a 30◦ undertwisted DNA
sequence illustrated in Figure 6, we find β = 0.725. This value is similar to that found in the
shorter simulation of 10 ns sampled every 1 ps, where we found β = 0.750; the value of 0.725
is identical to the value obtained from the shorter AMBER simulation; both these values are
reported in Table 1.
For the 38◦ overtwisted DNA molecule the long-time SDE simulation gives β = 0.775, which
lies between the value of β = 0.875 from the shorter SDE simulation and β = 0.700 from the
12
shorter AMBER simulation. Thus for both twist angles, the long-time SDE simulation gives
exponents closer to the AMBER results than the shorter SDE simulations.
12
10
8
6
4
2
)
)
ω
(
T
F
D
(
g
o
l
0
−10
−8
−6
−4
log(ω)
−2
0
2
Figure 6: Log-log plot of the discrete Fourier transform (power spectrum) of y0(t) from a
long-time (100 ns) SDE simulation for a 30◦ twisted DNA sequence, y0(t) data sampled every
1 ps.
3.4 Bases away from the defect
Finally we have analysed the motion of the base-pairs adjacent and further away from the
defect in the AMBER and SDE systems, in both cases using the example of a 38◦ overtwisted
DNA helix. For example, Figure 8 illustrates the power spectrum of the second-neighbour
base-pair y2t, sampled every 1 ps over a simulation of length 10ns. Although the decay with
increasing frequency (ω) is not as clear as in Figures 3, 4, 5, 6, or 7, it is still possible to fit
a straight line through the points. For the AMBER and SDE simulations respectively, the
gradients of these lines are 0.225 and 0.210 respectively.
This procedure has been repeated for the first neighbour base-pair, y1(t), and more distant
base-pairs, y3(t), y4(t), and the corresponding gradients of the best fit lines of the log-log
plots have been calculated. Assuming that the FTs have power-law forms, these gradients,
which correspond to the exponents, β, are given in Table 3. From the data in this Table we
observe that β decreases as one moves away from the defect site. This behaviour is due to the
reduced influence of the breathing pair on the neighbouring base-pairs. We observe a drop
from β ≈ 1 at the defect (n = 0) to just under one quarter at the nearest neighbour (n = 1)
in both SDE and AMBER.
The analysis of y0(t) showed that the exponent β increased from 0.7 − 0.8 to 0.90 − 0.95
when the sampling frequency was decreased from 1ps to 2 fs, suggesting convergence to β = 1
13
12
10
8
6
4
2
)
)
ω
(
T
F
D
(
g
o
l
0
−10
−8
−6
−4
log(ω)
−2
0
2
Figure 7: Log-log plot of the discrete Fourier transform (power spectrum) of y0(t) from a
long-time (100 ns) SDE simulation for a 38◦ twisted DNA sequence, y0(t) data sampled every
1 ps.
Base-pair βAM BER
y1(t)
y2(t)
y3(t)
y4(t)
0.225
0.180
0.180
0.180
βSDE
0.210
0.135
0.130
0.130
Table 3: The gradient β of the log-DFT function, from AMBER and SDE data sampled every
1 ps for 10 ns.
Base-pair βAM BER
y1(t)
y2(t)
y3(t)
y4(t)
0.450
0.425
0.410
0.390
βSDE
0.445
0.205
0.180
0.155
Table 4: The gradient β of the log-DFT function, from AMBER and SDE data sampled every
2 fs for 2 ns.
in the limit of small sampling frequency. Hence, we attempt to find more accurate values
for the β-exponent for the neighbouring bases by decreasing the sampling timestep from 1
ps to 2 fs (as we reduce the simulation length from 10ns to 2ns). We obtain the gradients
given in Table 4. Summarising, we find values just under one half at the nearest neighbour
(n = 1) in both SDE and AMBER. For y2, in AMBER, there is then a further slow decay of
14
)
)
ω
(
T
F
D
(
g
o
l
4
2
0
−2
−8
(a)
−6
−4
−2
log(ω)
(b)
0
2
)
)
ω
(
T
F
D
(
g
o
l
4
2
0
−2
−8
−6
−4
−2
log(ω)
0
2
Figure 8: Discrete Fourier transforms (power spectra) of y2(t) plotted against ω on a log-log
scale. Data sampled every 1 ps for 10 ns from a simulation of a 38◦ twisted DNA sequence
from (a) AMBER, and (b) the SDE system (2) -- (5).
O(0.01) per base-pair; whereas in SDE there is a more significant drop to 0.2 for y2 and then
a slow decrease of O(0.01) per base-pair. We observe once again that these later results with
a decreased sampling timestep of ∆t = 2 fs increases the measured values of β. Hence, we
recommend that the even with spatially coarse-grained models of DNA, kinetic characteristics
should be analysed using as small a timestep as possible, even as small as 2 fs (as used in
AMBER) in order to obtain correct results and conclusions.
This extra decrease in the SDE system may be due to the reduced number of degrees of
freedom (only one per base-pair in the SDE), whereas in AMBER there are O(102) degrees
of freedom per base-pair (the bases having 15 atoms moving in 3D space, in addition to the
phosphate backbone). Whilst all base-pairs receive energy in the form of white noise forcing
(in the SDE system) and in the form of random collisions with water molecules (AMBER),
this energy is used and dissipated differently in the defective base from its neighbours. At
the defect, there is a change in temporal behaviour, since white input noise (ξ0) is converted
to pink output (y0) giving β ≈ 1, whereas in the neighbouring bases, the output noise (yn)
remains significantly closer to white, that is β is significantly smaller.
15
4 Conclusions
In summary, we have simulated a sequence of 12 base-pairs of DNA using two different models
for a variety of times up to 100ns. We have analysed the displacement between the defective
base-pair near the centre of the chain. Fourier transforms of the distance trajectory taken
from both the AMBER and the mesoscopic stochastic differential equation simulations exhibit
scale-free, or critical, behaviour for all twist angles in the range 30◦-40◦ per base-pair. From
(6), we have by0(ω) = Cω−β across a considerable range of frequencies, ω. Furthermore, we
find that β = 1 for all twist angles. Although we have imposed white noise forcing in the
system (2) -- (5), the noise observed in the output y0(t) is pink (β ≈ 1, Figure 3). This shows
that our SDE system preserves the SOC properties of DNA observed in the fully deterministic
all-atom AMBER simulations.
It is suspected that proteins which interact with DNA overtwisting or undertwisting the
structure in order to ease the release of bases out from the structure. The fact that β = 1
for all twist angles appears to suggest that such a strategy will not change the base-pair
breathing. However, the constant C in the formula by0(ω) = Cω−β will depend on twist angle,
and mean that the fraction of time spent in the breathing state is less for the more stable
angles 35◦ − 36◦ and more for the overtwisted or undertwisted DNA structures (i.e those in
the ranges 38◦ − 40◦ and 30◦ − 34◦ respectively).
What is surprising about our results is that the critical behaviour is not specific to any one
twist angle but occurs at all angles. One might expect that, at normal twist angles of 35◦ -- 36◦,
stable behaviour would be observed, with breathing events having some short characteristic
timescale; at smaller and larger twist angles, a critical point would be found, where the DNA
exhibited scale-invariant breathing, and that at even more extreme twist angles, the open state
would be stable. However, this is not the case at all, instead, we find 1/f behaviour at all
twist angles. Since the emergence of this critical behaviour is not affected by the variation of
the twist angle of the system parameters values, or by the careful tuning of other parameters,
we describe this as self-organised criticality.
The scale-free nature of the kinetics of breathing events at all twist angles described herein is
strongly reminiscent of the behaviour of fluctuations in systems at criticality. Thus, it appears
that without any tuning of the interaction parameters of the DNA strand, it is at a critical
point where open bubbles spontaneously nucleate, hence we apply the term 'self-organized
criticality'. Figures 3 and 7 suggest that there is an upper frequency (around loge ω = −1),
above which the amplitude of base-pair separation modes is small but ceases to decay any
further, due to the effect of noise in the system. This cutoff is not due to the start of the
phonon band (which occupies the range √γ < ω < √γ + 4k), and which corresponds to a
relatively narrow range of velocities around 11 < ω < 14 (precise values depend on the twist
angle).
We observe some artifacts of the phonon band in the region of log ω being between two and
three and the defect mode near log ω = 1, in that there is a shoulder in the power spectrum
in Figure 5 where the trace is slightly larger than expected; however, no behaviour should
be expected to persist over all scales, and the figures show good agreement with scale-free
kinetic behaviour (straight-line) over the considerably large range of −6 < log ω < 1.
One might think that the defect site is the cause of the SOC breathing behaviour. However,
16
replacing the thymine (T) base with a difluorotoluene (F) base lowers the barrier between the
closed and open states. The replacement does not affect the DNA structure or other behaviour,
as discussed in several papers, for example, (37). Lowering the energy barrier allows breathing
to occur at lower energies, and so occur with a higher frequency, on a timescale accessible
to MD simulations (on the nanosecond timescale as opposed to the microsecond scale for a
normal DNA sequence). There is no reason to suppose that a change in the frequency of
events should cause a more significant change in qualitative behaviour. Hence, we speculate
that in pure DNA, with no defect, but with multiwelled potentials between all corresponding
base-pairs, curves such as that seen in Figure 3, will be repeated but that the crossover
frequency (from ω-independent noise to breathing with amplitude proportional to 1/ω) will
be shifted to much lower frequencies, namely the microsecond scale, which is beyond current
MD simulations. Here we only have a double-well potential at the defect, the other inter-base
interactions are all governed by harmonic potentials, in reality, all inter-base interactions are
double-welled, this will allow an open base-pair to be the nucleus for a bubble of several
consecutive open base-pairs to form, as the along-chain interactions would then ease the
opening of neighbouring base-pairs.
Acknowledgments
CID was funded by the EU as part of MMBNOTT -- an Early Training Research Programme
in Mathematical Medicine and Biology.
References
[1] D. Hennig - Formation and propagation of oscillating bubbles in DNA initiated by structural distortions,
Eur. Phys. J. B 37, 391 -- 397, (2004).
[2] T. Ambjornsson, S.K. Banik, M.A. Lomholt & R. Metzler - Master equation approach to DNA breathing
in heteropolymer DNA, Phys Rev E 75, 021908, (2007).
[3] D.A. Case, T.A. Darden, T.E. Cheatham, III, C.L. Simmerling, J. Wang, R.E. Duke, R. Luo, K.M. Merz,
D.A. Pearlman, M. Crowley, R.C. Walker, W. Zhang, B. Wang, S. Hayik, A. Roitberg, G. Seabra,
K.F. Wong, F. Paesani, X. Wu, S. Brozell, V. Tsui, H. Gohlke, L. Yang, C. Tan, J. Mongan, V. Hornak,
G. Cui, P. Beroza, D.H. Mathews, C. Schafmeister, W.S. Ross, & P.A. Kollman, AMBER 9, University
of California, San Francisco (2006).
[4] C.I. Duduiala, J.A.D. Wattis, I.L. Dryden & C.A. Laughton - Nonlinear breathing modes at a defect site
in DNA, Phys Rev E, 80, 061906, (2009).
[5] J.A.D. Wattis, S.A. Harris, C.R. Grindon & C.A. Laughton - Dynamic model of base pair breathing in a
DNA chain with a defect, Phys Rev E, 63, 061903, (2001).
[6] J.A.D. Wattis - Nonlinear breathing modes due to a defect in a DNA chain, Phil Trans Roy Soc Lond A,
362, 1461 -- 1477, (2004).
[7] J.D. Watson & F.H.C. Crick - Molecular structure of nucleic acids, Nature 171, 737 (1953).
[8] E. Cubero, E.C. Sherer, F.J. Luque, M. Orozco & C.A. Laughton - Observation of spontaneous base
pair breathing events in the molecular dynamics simulation of a difluorotoluene-containing DNA oligonu-
cleotide, J. Am. Chem. Soc. 121, 8653 -- 8654, (1999).
[9] M. Peyrard & A.R. Bishop - Statistical mechanics of a nonlinear model for DNA denaturation, Phys
Rev, 62, 2755-2758 (1989).
[10] M. Peyrard & J. Farago - Nonlinear localization in thermalized lattices: Application to DNA, Physica A
288, 199-217 (2000).
[11] M. Peyrard, S.C. L´opez & G. James - Modelling DNA at the mesoscale: a challenge for nonlinear
science?, Nonlinearity 21, T91-T100 (2008).
17
[12] G. Gaeta & L. Venier - Solitary waves in twist-opening models of DNA dynamics, Phys Rev E 78, 011901
(2008).
[13] G. Caldarelli, R. Frondoni, A. Gabrielli1, M. Montuori1, R. Retzlaff & C. Ricotta - Percolation in real
wildfires, Europhys Lett 56, 510-516 (2001).
[14] M. Cadoni, R. De Leo & G. Gaeta - A composite model for DNA torsion dynamics, Phys Rev E 75
021919 (2007).
[15] D. Ruelle - Small random perturbations of dynamical systems and the definition of attractors, Commun
Math Phys 82, 137-151 (1981).
[16] S. Wolfram - Cellular automata as model of complexity, Nature 311, 419 (1984).
[17] H. Haken - Cooperative phenomena in systems far from thermal equilibrium and in nonphysical systems,
Rev Mod Phys 47, 67-121 (1975).
[18] J. Zinn-Justin - Quantum Field Theory and Critical Phenomena, Oxford University Press (2002).
[19] B. Chopard & M. Droz - Cellular Automata Modeling of Physical Systems, Cambridge University Press
(1998).
[20] G. Peng & D. Tian - The fractal nature of a fracture surface, J Phys A: Math Gen 23, 3257-3261 (1990).
[21] M.E.J. Newman - Power laws, Pareto distributions and Zipf 's law, Contemporary Physics 46(5), 323-351
(2005).
[22] P. Bak, C. Tang & K. Wiesenfeld - Self-organized criticality: An explanation of the 1/f noise, Phys Rev
Lett 59, 381 -- 384, (1987).
[23] P. Bak, C. Tang & K. Wiesenfeld - Self-organized criticality, Phys Rev A 38, 364 -- 374, (1988).
[24] S.C. Chapman - Inverse cascade avalanche model with limit cycle exhibiting period doubling, intermit-
tency, and self-similarity, Phys Rev E 62, 1905 -- 1911, (2000).
[25] Z. Olami, H.J.S. Feder & K. Christensen- Self-organised criticality in a continuous, nonconservative
cellular automaton modelling earthquakes, Phys Rev Lett 68, 1244 -- 1247, (1992).
[26] P. Bak, K. Christensen, L. Danon & T. Scanlon - Unified scaling law for earthquakes, Phys Rev Lett 88,
178501, (2002).
[27] P. Bak, K. Chen & C. Tang - A forest-fire model and some thoughts on turbulence, Phys Lett A 147,
297 -- 300, (1990).
[28] B. Drossel & F. Schwabl - Self-organized critical forest-fire model, Phys Rev Lett 69, 1629 -- 1632, (1992).
[29] J.C. Phillips - Scaling and self-organized criticality in proteins I, Proc Natl Acad Sci 106(9), 3107 -- 3112,
(2009).
[30] J.C. Phillips - Scaling and self-organized criticality in proteins II, Proc Natl Acad Sci 106(9), 3113 -- 3118,
(2009).
[31] G. Werner - Metastability, criticality and phase transitions in brain and its models, Biosystems 90, 496 --
508, (2007).
[32] A.M. Selvam - Universal spectrum for DNA base C-G frequency distribution in Human chromosomes 1
to 24, arXiv:physics/0701079 (2007).
[33] S.A. Harris, E. Gavathiotis, M.S. Searle, M. Orozco & C.A. Laughton - Cooperativity in drug-DNA
recognition: a molecular dynamics study, J Am Chem Soc 123, 12658 -- 12663, (2001).
[34] J. Schlitter - Estimation of absolute and relative entropies of macromolecules using the covariance matrix,
Chem. Phys. Lett. 215, 617 -- 621, (1993).
[35] C.I.
Duduiala
-
Stochastic Nonlinear Models
of DNA Breathing
at
a Defect,
http://etheses.nottingham.ac.uk/, PhD Thesis, University of Nottingham, (2009).
[36] M. Peyrard, S. C. L´opez, D. Angelov - Fluctuations in the DNA double helix, Eur Phys J Special Topics
147, 173-189 (2007).
[37] K.M. Guckian, T.R. Krugh & E.T. Kool - Solution structure of a DNA duplex containing a replicable
difluorotoluene-adenine pair, Nature Structural Biology 5, 954 -- 959, (1998).
18
|
1806.10490 | 1 | 1806 | 2018-06-27T13:58:02 | Statistical mechanics of an elastically pinned membrane: Equilibrium dynamics and power spectrum | [
"physics.bio-ph",
"cond-mat.soft"
] | In biological settings membranes typically interact locally with other membranes or the extracellular matrix in the exterior, as well as with internal cellular structures such as the cytoskeleton. Characterization of the dynamic properties of such interactions presents a difficult task. Significant progress has been achieved through simulations and experiments, yet analytical progress in modelling pinned membranes has been impeded by the complexity of governing equations. Here we circumvent these difficulties by calculating analytically the time-dependent Green's function of the operator governing the dynamics of an elastically pinned membrane in a hydrodynamic surrounding and subject to external forces. This enables us to calculate the equilibrium power spectral density for an overdamped membrane pinned by an elastic, permanently-attached spring subject to thermal excitations. By considering the effects of the finite experimental resolution on the measured spectra, we show that the elasticity of the pinning can be extracted from the experimentally measured spectrum. Membrane fluctuations can thus be used as a tool to probe mechanical properties of the underlying structures. Such a tool may be particularly relevant in the context of cell mechanics, where the elasticity of the membrane's attachment to the cytoskeleton could be measured. | physics.bio-ph | physics | Statistical mechanics of an elastically pinned membrane: Equilibrium dynamics and
power spectrum
Josip A. Janes,1, 2 Daniel Schmidt,1, 3 Udo Seifert,3 and Ana-Suncana Smith1, 2, ∗
1Institut fur Theoretische Physik and Cluster of Excellence: Engineering of Advanced Materials,
Friedrich Alexander Universitat Erlangen-Nurnberg, 91052 Erlangen, Germany
3II. Institut fur Theoretische Physik, Universitat Stuttgart, 70569 Stuttgart, Germany
2Institut Ruder Boskovi´c, 10000 Zagreb, Croatia
In biological settings membranes typically interact locally with other membranes or the extra-
cellular matrix in the exterior, as well as with internal cellular structures such as the cytoskeleton.
Characterization of the dynamic properties of such interactions presents a difficult task. Signifi-
cant progress has been achieved through simulations and experiments, yet analytical progress in
modelling pinned membranes has been impeded by the complexity of governing equations. Here we
circumvent these difficulties by calculating analytically the time-dependent Green's function of the
operator governing the dynamics of an elastically pinned membrane in a hydrodynamic surrounding
and subject to external forces. This enables us to calculate the equilibrium power spectral density
for an overdamped membrane pinned by an elastic, permanently-attached spring subject to thermal
excitations. By considering the effects of the finite experimental resolution on the measured spec-
tra, we show that the elasticity of the pinning can be extracted from the experimentally measured
spectrum. Membrane fluctuations can thus be used as a tool to probe mechanical properties of the
underlying structures. Such a tool may be particularly relevant in the context of cell mechanics,
where the elasticity of the membrane's attachment to the cytoskeleton could be measured.
I.
INTRODUCTION
A phospholipid membrane can be easily deformed and
exhibits appreciable fluctuations due to its small elas-
tic constant [1–6]. While occurring on time scales be-
tween 10−9 and 10−5 s [7–10], the fluctuations are over-
damped by the surrounding fluid [11–14]. Nonethe-
less, mean fluctuation amplitudes of up to 100 nm have
been observed experimentally [15].
In a vicinity of a
substrate, these fluctuations are known to contribute
to an effective potential which prevents the membrane
from non-specifically adhering to the underlying scaffold
[1, 3, 10, 16–21]. The scaffold in turn affects the hydro-
dynamic damping of the membrane reflected in changes
of the time dependent correlation function and the asso-
ciated power spectrum, the so-called power spectral den-
sity (PSD) [15, 22–24]. Moreover, in the cellular environ-
ment, active processes couple with the membrane fluctua-
tions [23–33], resulting in the violation of the fluctuation-
dissipation theorem in the activated state of the cell [34–
36].
Over the last two decades, models for the fluctuations
of free membranes, based on the Helfrich energetics and
Stokes fluid dynamics, were experimentally confirmed,
either by measuring the PSD or its Fourier transform,
the time dependent correlation function (for reviews see
[7, 15] and references therein). This body of work con-
firmed the appropriateness of the models based on the
Helfrich Hamiltonian [1] to capture the equilibrium dy-
namics of free membranes. However, the presence of the
∗ author
to whom correspondence
should be
addressed:
[email protected]
pinning introduces a challenge, which is often circum-
vented by homogenising the effects of interactions with
the scaffold [36, 37]. Alternative approaches, where the
local pinnings remain explicit, commonly involved sim-
ulations [38–41], while the theoretical modelling focused
on the static properties of the membrane shape and fluc-
tuations [37, 42, 43]. On the other hand, analytic treat-
ments of the membrane-dynamics, even in the context of
equilibrium, remained an open problem up to now. First
quasi-analytical predictions were obtained for infinitely
strong pinnings [44], while the more realistic case, where
the membrane is attached by proteins which themselves
maintain a certain flexibility has not been considered so
far.
In this paper we address this open issue by analyti-
cally solving the integro-differential equation governing
the motion of a pre-tensed membrane pinned by a sin-
gle flexible construct. We describe in detail the effect
of the pinning on the membrane's equilibrium dynam-
ics. At last, we provide an exact analytical method for
calculating the pinning stiffness from the experimentally
measured PSD, accounting for the finite resolution of
the setup. While constructed in the context of biolog-
ical membranes, the obtained result can be applied more
generally, in the context of bending fluctuations of thin
sheets.
II. EQUATION OF MOTION
The system consists of one flexible attachment (har-
monic spring of an elastic constant λ and rest length l0)
that pins a tensed membrane (bending rigidity κ, tension
σ). The later fluctuates in a harmonic non-specific po-
tential (strength γ) positioned at the distance h0 above
8
1
0
2
n
u
J
7
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
0
9
4
0
1
.
6
0
8
1
:
v
i
X
r
a
2
Figure 1. Snapshot from the Langevin simulation of locally pinned membrane fluctuating in a non-specific potential (Left).
Sketch of the system (Right). Membrane is residing in a harmonic potential of strength γ at h0 separation from a flat substrate
and pinned by an elastic spring of stiffness λ and rest length l0 positioned at r = 0.
the substrate (Fig. 1). Placing the origin of the coor-
dinate system at the pinning site and into the minimum
of the non-specific potential sets the form of the energy
functional [43, 45] as
(cid:90)
(cid:34)
H =
dr
κ
2
(cid:0)∇2u(r, t)(cid:1)2
(∇u(r, t))2 +
+
σ
2
γ
2
(u(r, t))2
(cid:35)
A
λ (u(r, t) − (l0 − h0))2 δ(r)
+
1
2
.
(1)
Here and throughout the paper, the energy scale kBT is
set to unity, with Boltzmann constant denoted as kB and
absolute temperature T .
Dynamics of an overdamped membrane in a hydrody-
namic surrounding is captured by the Langevin equation
[20, 39, 41, 46–50]
∂u(r, t)
∂t
=
(cid:90)
(cid:18)
dr(cid:48)Λ(r − r(cid:48))
− δH
(cid:18)
δu(r, t)
≡ Λ(r) ∗
(cid:19)
− δH
δu(r(cid:48), t)
+ f (r(cid:48), t)
(cid:19)
+ f (r, t)
,
(2)
which states that the velocity of the membrane profile
u(r, t) is given by a convolution of the hydrodynamic ker-
nel Λ(r), the Oseen tensor, with the forces acting on the
membrane. External forces on the system are denoted by
f (r, t), while the internal forces acting to minimize the
Hamiltonian (eq. (1)) are given by the first variation [43]
=(cid:0)κ∇4 − σ∇2 + γ + λδ(r)(cid:1) u(r, t)
δH
δu(r, t)
−λ(l0 − h0)δ(r).
(3)
Together with eq. (2), this leads to the equation for the
dynamics of an overdamped pinned membrane
D [u(r, t)] = Λ(r) ∗ [f (r, t) + λ(l0 − h0)δ(r)] ,
(4)
with the operator D set as
+ Λ(r) ∗(cid:0)κ∇4 − σ∇2 + γ + λδ(r)(cid:1) .
D =
∂
∂t
where Λk is the spatial Fourier transform of Λ(r) and
(5)
Ek = κk4 + σk2 + γ.
(11)
III. TIME-DEPENDENT GREEN'S FUNCTION
The solution of eq. (4) provides the evolution of the
membrane profile u(r, t). It is obtained by the integration
of forces acting on the membrane, the latter accounted
for by the dynamic Green's function g(r, tr(cid:48), t(cid:48))
(cid:90)
dr(cid:48)(cid:90)
u(r, t) =
(cid:48)
dt
g(r, tr(cid:48), t(cid:48))×
R
R2
× (f (r(cid:48), t(cid:48)) + λ(l0 − h0)δ(r(cid:48))) .
(6)
(7)
Here the Green's function is defined by
D [g(r, tr(cid:48), t(cid:48))] =Λ(r) ∗ [δ(r − r(cid:48))δ(t − t(cid:48))] .
Besides imposing causality, this equation is subject to ho-
mogeneous spatial boundary conditions forcing the mem-
brane in the minimum of the non-specific potential far
from the pinning.
A. Free membrane (λ = 0)
Recognizing spatio-temporal translational invariance
of the free membrane system, the corresponding Green's
function gf (r, tr(cid:48), t(cid:48)) can be written in terms of variables
t = t − t(cid:48) and r = r − r(cid:48) as
gf (r, tr(cid:48), t(cid:48)) = gf (r − r(cid:48), t − t(cid:48)) ≡ gf (r, t).
(8)
Consequently, for the free membrane eq. (7) becomes
(cid:20) ∂
+ Λ(r) ∗(cid:0)κ∇4
= Λ(r) ∗(cid:2)δ(r)δ(t)(cid:3) .
∂t
r + γ(cid:1)(cid:21)
r − σ∇2
gf (r, t)
(9)
Fourier transforming eq. (9) (r → k and t → ω) upon
rearranging yields
gf (k, ω) =
1
iω/Λk + Ek
,
(10)
u(r)h0γu2(r)l0ru(r,t)r=0λFinally, transforming back to the spatio-temporal domain
(k → r and ω → t) provides the spatio-temporal Green's
function for the free membrane
(cid:90)
(cid:90)
dω
2π
gf (r, t) =
dk
eikreiωt
R2
(2π)2
R
iω/Λk + Ek
.
(12)
Integrating over the frequencies gives
(cid:90)
(cid:90) ∞
(2π)2 Λkeikre−ΛkEk tΘ(t)
dk
gf (r, t) =
=
dk
2π
0
ΛkkJ0(kr)e−ΛkEk tΘ(t),
(13)
where Θ is the Heaviside step function appearing as a
consequence of causality. Moreover, gf (r, t) depends only
on the absolute value of r, as expected.
For ω = 0, the Green's function gf (r, ω) reduces to the
static correlation function [43]. Consequently,
(cid:32)(cid:114)(cid:16) λ0
(cid:114)(cid:16) λ0
(cid:33)
(cid:17)2 − 1
(cid:17)2 − 1
m
4σ
m
4σ
arctan
2πσ
3
For r = 0 in eq. (19) we find
g(r, ωr(cid:48)) =
gf (r(cid:48), ω)
1 + λgf (r = 0, ω)
,
(20)
which upon inserting into (19) yields the Green's function
for the pinned membrane in the spatio-frequency domain
g(r, ωr(cid:48)) = gf (r − r(cid:48), ω) − λ
gf (r, ω)gf (r(cid:48), ω)
1 + λgf (r = 0, ω)
.
(21)
Fourier transforming eq. (21) (r → k and r(cid:48) → k(cid:48)) results
in
g(k, ωk(cid:48)) = (2π)2δ(k + k(cid:48))gf (k, ω) − λ
gf (k, ω)gf (k(cid:48), ω)
1 + λgf (r = 0, ω)
(22)
,
which is the representation of the Green's function in the
Fourier space.
gf (r = 0, ω = 0) ≡ 1
λm
=
(14)
IV. DYNAMICS OF THERMAL
FLUCTUATIONS
represents the fluctuation amplitude, which in the ten-
sionless case reduces to
A. Oseen tensor
gf (r = 0, ω = 0; σ = 0) ≡ 1
λ0
m
=
1
√
κγ
8
.
(15)
In thermal equilibrium f (r, t) is associated with the
stochastic thermal noise characterized by a vanishing
mean
B. Pinned membrane
Permanent pinning breaks the spatial, but keeps the
temporal translational invariance. Therefore, the Green's
function of the pinned membrane must be described by
two spatial variables r and r(cid:48) and one temporal variable
t = t − t(cid:48)
g(r, tr(cid:48), t(cid:48)) = g(r, t − t(cid:48)r(cid:48)) ≡ g(r, tr(cid:48)).
(16)
(cid:20) ∂
In this notation, eq. (7) for the pinned membrane Green's
function becomes
+ Λ(r) ∗(cid:0)κ∇4 − σ∇2 + γ + λδ(r)(cid:1)(cid:21)
= Λ(r) ∗(cid:2)δ(r − r(cid:48))δ(t)(cid:3) . (17)
g(r, tr(cid:48))
∂t
Fourier transforming (r → k and t → ω) and rearranging
eq. (17) gives
g(k, ωr(cid:48)) =
e−ikr(cid:48)
iω/Λk + Ek
− λg(r = 0, ωr(cid:48))
1
iω/Λk + Ek
.
(18)
Transforming back to the spatial domain (k → r) gives
g(r, ωr(cid:48)) =gf (r − r(cid:48), ω) − λg(r = 0, ωr(cid:48))gf (r, ω). (19)
(cid:104)f (r, t)(cid:105) = 0
(23)
and spatio-temporal correlations obeying the fluctuation-
dissipation theorem
(cid:104)f (r, t)f (r(cid:48), t(cid:48))(cid:105) = 2Λ−1(r − r(cid:48))δ(t − t(cid:48)).
(24)
Here Λ−1(r) is defined by
Λ(r) ∗ Λ−1(r) = δ(r).
(25)
To model damping of the membrane due to hydrody-
namic interactions with the surrounding fluid close to a
wall [46, 48], we will use the Fourier transform of the
Oseen tensor Λk
Λk = (4ηk)−1,
(26)
where η is the viscosity of the surrounding fluid. Eq.
(26) is appropriate when the wall is permeable to the
fluid. In the presence of an impermeable wall, damping
coefficients are modified [51, 52], which in the case of
protein mediated adhesion, typically has an effect only
on the amplitude of the first few membrane modes [49].
Furthermore, if the membrane is surrounded by two dif-
ferent fluids with viscosities η1 and η2, the viscosity η
in the damping coefficients is replaced by the arithmetic
mean η = (η1 + η2)/2 [15].
B. Simulation methods
Eq. (4) for the membrane dynamics, subject to ther-
mal noise defined with eqs. (23)-(26), is the foundation
of our Langevin dynamics simulations of the membrane,
described previously in full detail [41].
In the current
case, one pinning site is placed in a middle of the sim-
ulation box (periodic boundary conditions) of a size of
640 × 640 nm for a tensionless membrane, and a size of
5120 × 5120 nm at finite tensions. The simulations are
performed with a temporal and lateral resolution of 10−9
s and 10 nm, respectively. The membrane height profile
is recorded as a function of time and analyzed to extract
the membrane shape and correlation functions.
C. Power Spectral Density
We complement the simulations of thermally fluctu-
ating membrane with the analytic calculations based on
the Green's function approach (eq. (6)). We start with
rewriting eq. (6) as
u(r, t) = (cid:104)u(r)(cid:105) + v(r, t),
where
v(r, t) =
(cid:48)
dt
g(r, t − t(cid:48)r(cid:48))f (r(cid:48), t(cid:48))
(28)
are the fluctuations around the ensemble averaged static
profile [43]
(cid:104)u(r)(cid:105) =
(cid:48)
dt
g(r, t − t(cid:48)r(cid:48))λ(l0 − h0)δ(r(cid:48)).
(29)
Transforming (t → ω) eq. (28) gives
(cid:90)
R2
R
dr(cid:48)(cid:90)
dr(cid:48)(cid:90)
R
(cid:90)
R2
v(r, ω) =
dr(cid:48) g(r, ωr(cid:48))f (r(cid:48), ω)
(30)
=
(cid:90)
R2
(2π)2
×
1
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)1 −
∞(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)1 −
0
dk
from which the PSD (cid:104)v(r, ω)2(cid:105) can be calculated as (see
Supplementary Information)
(cid:104)v(r, ω)2(cid:105) =
2Λ−1
k
×
dk
R2
(ω/Λk)2 + E2
k
λgf (r, ω)
e−ikr
1 + λgf (r = 0, ω)
(cid:12)(cid:12)(cid:12)(cid:12)2
.
(31)
With the hydrodynamic coefficients specified as in eq.
(26), eq. (31) becomes
(cid:104)v(r, ω)2(cid:105) =
4η
π
×
(4ηkω)2 + (κk4 + σk2 + γ)2×
k2
λgf (r, ω)
e−ikr
.
(32)
1 + λgf (r = 0, ω)
(cid:12)(cid:12)(cid:12)(cid:12)2
4
In the absence of the pinning (λ = 0), eq. (32) becomes
homogeneous in space and reduces to the well-known re-
sult(cid:10)vf (ω)2(cid:11) =
4η
π
∞(cid:90)
dk
k2
(4ηkω)2 + (κk4 + σk2 + γ)2 ,
0
(33)
which for small and large ω has the limiting behaviour
[9, 11, 23, 53]
(cid:104)vf (ω)2(cid:105) =
η√
√
γ(λ0
m/4+σ)3
, ω (cid:28) ω0
ω−5/3, ω (cid:29) ω0.
(34)
Here, ω0 ≡ 4(cid:112)κγ3/η is a cross-over frequency, defined as
1
2η2κ
6 3
the intersection of the lines fitting the low- and high-
frequency limits of the spectrum. The low-frequency
limit decays with σ−3/2 for tensions σ (cid:29) λ0
m (Fig. 2b).
It is clear from Eq. (32) that the PSD at the pinning
site r = 0 can be recast into
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)2 (cid:104)vf (ω)2(cid:105). (35)
(27)
(cid:104)v(r = 0, ω)2(cid:105) =
1
1 + λgf (r = 0, ω)
In agreement with simulations based on eqs. (4), (23)
and (24) [41], eq. (35) shows that only the pinning stiff-
ness, and not its length, has an effect on the PSD and
that the pinning affects only the low-frequency regime
(Fig. 2a). The low-frequency behaviour can be obtained
upon combining eq. (35) for ω = 0 with eq. (34) to yield
(cid:18)
(cid:19)2
(cid:112)γ(λ0
η
m/4 + σ)3
4σ/λ0
4σ/λ0
m (cid:28) 1
m (cid:29) 1.
(36)
(cid:16)
(cid:16)
(cid:104)v(r = 0, ω = 0)2(cid:105) =
1
1 + λ/λm
(cid:17)2
1
1+λ/λ0
m
1
1+λ ln(σ)/(2πσ)
η√
γ(λ0
(cid:17)2
,
m/4)3
η√
,
γσ3
Eq. (36) shows that the low-frequency spectrum is in-
m (cid:28) 1 and it
dependent of membrane tension for 4σ/λ0
decays with σ−3/2 for membrane tensions large enough
to diminish the effect of the pinning (λ ln(σ)/(2πσ) (cid:28) 1)
2b). For stiff pinnings (λ/λm (cid:29) 1) the low-
(Fig.
frequency limit falls off as λ−2 (Fig. 2c). On the other
hand, for λ/λm (cid:28) 1, pinning effects vanish even in
the low-frequency limit. Interestingly, the cross-over fre-
quency ω0 for the pinned membrane
(cid:104)
6 3(cid:112)
2η2κ (cid:104)v(r = 0, ω = 0)2(cid:105)(cid:105)−3/5
,
(37)
ω0 =
defined analogously to the one for the free membrane,
becomes sensitive to the elastic properties of the pinning
as ω0 ∼ λ6/5 and as such increases with the pinning stiff-
ness.
5
Figure 2. Dynamical properties of a membrane at the pinning site. Comparison of modelling (lines) and simulations (symbols)
shows excellent agreement across the entire parameter range. a) Power spectral density of a free membrane (eq. (33)) (red) and
a pinned tensionless membrane eq. (35) (blue dashed curves for λ/λ0
m = 1, 3, 10, increasing in the direction of the arrows). The
high-frequency regime of the PSD is unaffected by the pinning and the free-membrane behaviour (ω−5/3) is recovered. b) Low
frequency limit of the PSD (eq. (36)) as a function of the membrane tension σ for different pinning strengths (λ/λ0
m = 1, 3, 10).
For large tensions, a σ−3/2 dependence is recovered irrespective of λ. c) Low frequency limit as a function of the pinning
m = 1, 3, 10). For large bond stiffness, a λ−2 dependence is displayed. All
strength for different membrane tensions (8σ/λ0
curves are plotted for κ = 20kBT , γ = 3 × 10−7kBT /nm4 and η = 1 mPas.
V. EFFECT OF THE FINITE EXPERIMENTAL
section I. A.1.)
RESOLUTION ON THE FLUCTUATION
SPECTRUM
In order to compare with experiments, it is necessary to
account for the finite temporal and spatial resolutions of
the set-up [15, 54]. Averaging the true membrane profile
u(r, t) over a spatial domain A and a time interval τ , gives
rise the so-called apparent membrane profile uA
τ (r, t)
τ
(cid:104)vR2π
dk
∞(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) J1(kR)
k
0
×
×
(r = 0, ω)2(cid:105) =
kΛ−1
k
(ω/Λk)2 + E2
k
4
R2
−
λ
1 + λgf (0, ω)
(cid:18) sin(ωτ /2)
(cid:19)2 1
ωτ /2
π
×
∞(cid:90)
0
1
2π
dk(cid:48)
J1(k(cid:48)R)
iω/Λk(cid:48) + Ek(cid:48)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2
.
(40)
τ(cid:90)
0
(cid:90)
A
dt(cid:48)
τ
dr(cid:48)
A
uA
τ (r, t) =
u(r + r(cid:48), t + t(cid:48)),
(38)
Eq. (40) reduces to eq. (35) in the limit τ, R → 0.
from which it is straightforward to derive the apparent
PSD (Supplementary Information, section I. A)
(cid:104)vA
(cid:18) sin(ωτ /2)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
τ (r, ω)2(cid:105) =
(cid:90)
(cid:18)
(ω/Λk)2 + E2
k
1 − λgf (r + r(cid:48), ω)
Λ−1
ωτ /2
dk
R2
k
1 + λgf (0, ω)
×
×
(cid:19)2
(cid:90)
A
2
(2π)2×
dr(cid:48)
A
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)2
eik(r+r(cid:48)) ×
e−ik(r+r(cid:48))
.
(39)
The PSD measured around the pinning placed centrally
in a circle of radius R is (Supplementary Information,
The high frequency regime of the averaged PSD re-
covers the averaging behaviour of the free membrane -
spatial averaging changes the decay from ω−5/3 to ω−2
as previously reported [9], while finite temporal resolu-
tion induces an additional attenuation of ω−2. Hence,
the PSD which is subject to both temporal and spatial
averaging decays as ω−4.
In the low frequency regime, temporal averaging plays
no role for ω < 1/τ , while the finite spatial resolution
has a more complex effect. Due to the interplay with the
effects of the pinning, the low frequency amplitude is not
a monotone function of the averaging area (Fig. 3). In-
creasing the averaging area up to some critical size (which
is approximately the area affected by the pinning) ampli-
fies the low-frequency components, but further increase
of the averaging area starts to attenuate them. This can
be understood as a competition of two effects; averaging
has the effect of attenuating the low frequency compo-
nents, as can be seen for the free membrane, but at the
same time, averaging reduces the effect of the pinning on
●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●10210410610-810-610-4ω[s-1]〈v(r=0,ω)2〉[nm2s]freepinnedfreeh0=l0h0≠l0λ/λm0~ω-5/301310a)●●●●●●●●●●●●●●●●●●10-210010210-610-510-410-38σ/λm0〈v(r=0,ω=0)2〉[nm2s]freepinnedfreepinnedλ/λm0~σ-3/201310b)●●●●●●●●●10-210010-810-710-610-5λ/λm0〈v(r=0,ω=0)2〉[nm2s]~λ-28σ/λm001310c)6
Figure 3. Effect of the finite resolution (averaging over a circle of radius R and a time interval τ ) on the PSD at the pinning (eq.
(40)). a) Averaging decreases the difference between the free and pinned PSD's, therefore reducing the effect of the pinning on
the PSD. Pinning has no effect on the high-frequency part of the spectrum, for both the averaged and the unaveraged spectrum.
b) The low-frequency part of the PSD is a non-monotone function of the spatial averaging area. On the other hand, increasing
the averaging area always attenuates the high-frequency components, which in this case fall as ∼ ω−2 instead of ∼ ω−5/3. c)
Time averaging introduces oscillatory behaviour of the PSD for which only the envelope of the PSD is shown. The combined
effect of the spatial and temporal averaging gives a ∼ ω−4 behaviour of the high-frequency regime . Parameters κ = 20kBT ,
σ = 10−20kBT /nm2 and γ = 3 × 10−7kBT /nm4, λ/λ0
m=10 and η = 1 mPas.
the PSD, which amplifies the low-frequency components.
Obviously, the later effect is stronger up to the critical
averaging area size, after which the first effect dominates.
Specifically, for ω = 0 we obtain (Supplementary Infor-
mation, section I.A.1)
(cid:104)vR2π
τ
=
4η
π
∞(cid:90)
0
dk
k2
E2
k
(r = 0, 0)2(cid:105) =
(cid:12)(cid:12)(cid:12)(cid:12) J1(kR)
kR/2
(cid:12)(cid:12)(cid:12)(cid:12)2
s(R)
−
λ
1 + λ/λm
,
(41)
where for brevity purposes we introduce a reduced coef-
ficient s(R)
s(R) =
×
with a±
×
√
R2π
1
σ − 4κγ
(cid:18) 1 − a−RK1(a−R)
− 1 − a+RK1(a+R)
1 ±
(cid:115)
(cid:18) λ0
(cid:19)21/2
σ
a± =
1 −
a2−
a2
+
.
(cid:19)
,
(42)
(43)
m
4σ
2κ
In order to calculate the pinning stiffness λ from the
PSD, eq. (41) can be inverted, which upon introduction
of coefficients L, a, b and c, yields
λ =
with
(cid:18) 1
L
,
(cid:19)−1
(cid:19)2 − c
a
− 1
λm
(cid:115)(cid:18) b
2a
L =
−
b
2a
,
m/4 + σ)3
dk
kJ1(kR)
E2
k
,
πs2(R)
4(cid:112)γ(λ0
∞(cid:90)
(cid:18) 2
(cid:19)2
∞(cid:90)
4s(R)
R
0
and
a =
b =
c =
R
0
dk
J 2
1 (kR)
E2
k
− π
4η
(cid:104)vR2π
τ
(r = 0, 0)2(cid:105).
(46)
The averaged spectrum is contained in the coefficient c.
When implemented numerically, eqs. (44-46) represent a
fast and exact method for obtaining information about
the pinning stiffness from the experimentally measured
PSD. Here we note that the deconvolution of the noise
associated with the experimental setting should be per-
formed prior to the extraction of the pinning stiffness.
VI. DISCUSSION AND CONCLUSION
Our calculation of the Green's function (eq. (21)) fully
resolves the dynamics of an overdamped, permanently
pinned membrane in a hydrodynamic surrounding (eq.
(2)). The solution is general in a sense that it works
for any forces acting on the membrane and enables one
to study the membrane dynamics in the presence of both
non-thermal and thermal perturbations. The later case is
resolved in this paper by the calculation of the thermal
equilibrium power spectral density (eq. (31)). For the
specific case of hydrodynamic damping close to a perme-
able wall, our analytical calculation (eq. (32)) is verified
with Langevin simulations in a broad range of param-
eters, such as the pinning stiffness, membrane tension,
(44)
(45)
10210410610-810-610-4ω[s-1]〈vτ=0A=R2π(r=0,ω=0)2〉[nm2s]freefreepinnedR=0nmpinnedR=20nmω-5/3ω-2a)10210410610-810-610-4ω[s-1]〈vτ=0A=R2π(r=0,ω=0)2〉[nm2s]freeR=0nmpinnedR=0nm50nm200nm500nmω-5/3ω-2b)10410610-1010-810-610-4ω[s-1]〈vτA=R2π(r=0,ω=0)2〉[nm2s]Rτ0nm0s0nm10-5s200nm0s200nm10-5sω-5/3ω-2ω-11/3ω-4c)7
and strength of the non-specific potential, which were al-
lowed to independently vary for several orders of magni-
tude (Fig. 2). It is shown that the pinning decreases the
low-frequency amplitudes of the spectrum and pushes the
cross-over frequency (eq. (37)) to higher values, while the
high-frequency amplitudes remain unaffected (Fig. 2).
Interestingly, the pinned-membrane PSD at the pin-
ning site is given by a product of a free-membrane PSD
and a λ-dependent prefactor (eq.
(35)). Assuming
knowledge of the pinning stiffness λ, this enables infer-
ence of the pinned-membrane PSD at the pinning site di-
rectly from the free-membrane PSD. This approach has a
clear advantage over a direct measurement of the pinned-
PSD, as it replaces the pinned-membrane measurement,
with a well-established free membrane measurement. On
the other hand, if λ is not known, it can be easily de-
termined by comparing the pinned- and free-membranes
PSDs. The low-frequency limit of the PSD at the pinning
site (eq. (36)) is particularly useful for getting a better
understanding of the interplay of the system parameters
and shows a λ−2 decay.
These relationships, however, may not be observed ex-
perimentally due to the finite resolution of the measure-
ments. Specifically, while the effects of the temporal aver-
aging are simple, significant spatial averaging introduces
nontrivial modulations of the spectrum and breaks the re-
lation between the free- and the pinned-membrane PSD
given by eq. (35).
In this regime the pinning stiffness
can be inferred from the measured PSD with the use of
the spatially averaged spectrum (eqs. (44-46)).
A deep understanding of the mechanics of the pinned
membrane is crucial for elucidating the role of more com-
plex pinnings, which under typical biological conditions
stochastically bind and unbind from the membrane. The
stochasticity of this attachment will have additional ef-
fects on fluctuations of the membrane which could not
be resolved prior to this investigation, and will be sub-
ject to a future study. The results presented in this pa-
per will thus help establishing the connection between
functioning of the protein assembly and the properties of
the elastic fluctuating membrane, which is important for
understanding of the formation of adhesions. Namely,
there is a growing body of evidence that the membrane
affects the affinity [55] and the kinetic rates for protein
binding [41, 56], which in turn affect the fluctuations and
the early stage signalling in developing junctions between
cells [57].
The model presented here can be used to measure the
elasticity of the bond from membrane fluctuations. Hith-
erto, it was not possible to interpret these measurements
accurately, a task that is enabled now by our current
work. Such measurements could then be compared to
AFM measurements, which are commonly used to study
elastic properties of the linkers.
Acknowledgments: A.-S.S and J.A.J. thank ERCStg
MembranesAct for support as well as the Croatian Sci-
ence Foundation research project CompSoLs MolFlex
8238. A.-S.S and D.S. were supported by the Re-
search Training Group 1962 at the Friedrich-Alexander-
Universitat Erlangen-Nurnberg.
[1] W. Helfrich, "Elastic properties of lipid bilayers: Theory
and possible experiments," Z. Naturforsch., C: J. Biosci.
28, 693–703 (1973).
[2] W. Helfrich, "Steric interaction of fluid membranes in
multilayer systems," Z. Naturforsch., A: Phys. Sci. 33,
305 (1978).
[3] R. Lipowsky and E. Sackmann, eds., Structure and dy-
namics of membranes (Elsevier, 1995).
[4] A.-S. Smith, B.G. Lorz, S. Goennenwein, and E. Sack-
mann, "Force-controlled equilibria of specific vesicle-
substrate adhesion," Biophys. J. 90, L52–L54 (2006).
[5] C. Monzel, S. F. Fenz, M. Giesen, R. Merkel, and K. Sen-
gupta, "Mapping fluctuations in biomembranes adhered
to micropatterns," Soft Matter 8, 6128–6138 (2012).
[6] S. Marx, J. Schilling, E. Sackmann, and R. Bruinsma,
"Helfrich repulsion and dynamical phase separation of
multicomponent lipid bilayers," Phys. Rev. Lett. 88,
138102 (2002).
[7] C. Monzel and K. Sengupta, "Measuring shape fluctua-
tions in biological membranes," J. Phys. D: Appl. Phys.
49, 243002 (2016).
[8] S. F. Fenz and K. Sengupta, "Giant vesicles as cell mod-
els," Integr. Biol. 4, 982–995 (2012).
[9] T. Betz and C. Sykes, "Time resolved membrane fluctu-
ation spectroscopy," Soft Matter 8, 5317–5326 (2012).
[10] D. Schmidt, C. Monzel, T. Bihr, R. Merkel, U. Seifert,
K. Sengupta, and A.-S. Smith, "Signature of a nonhar-
monic potential as revealed from a consistent shape and
fluctuation analysis of an adherent membrane," Phys.
Rev. X 4, 021023 (2014).
[11] F. Brochard and J. F. Lennon, "Frequency spectrum of
the flicker phenomenon in erythrocytes," J. Phys. France
36, 11 (1975).
[12] L. Kramer, "Theory of light scattering from fluctua-
tions of membranes and monolayers," J. Chem. Phys.
55, 2097–2105 (1971).
[13] U. Seifert and S.A. Langer, "Viscous modes of fluid bi-
layer membranes," Europhys. Lett. 23, 71 (1993).
[14] P. J. Atzberger, "Stochastic eulerian lagrangian meth-
ods for fluid–structure interactions with thermal fluctu-
ations," Journal of Computational Physics 230, 2821 –
2837 (2011).
[15] C. Monzel, D. Schmidt, U. Seifert, A.-S. Smith,
R. Merkel, and K. Sengupta, "Nanometric thermal fluc-
tuations of weakly confined biomembranes measured with
microsecond time-resolution," Soft Matter 12, 4755–4768
(2016).
[16] W. Helfrich and R.-M. Servuss, "Undulations, steric in-
teraction and cohesion of fluid membranes," Nuovo Ci-
mento D 3, 137–151 (1984).
[17] E. A. Evans and V. A. Parsegian, "Thermal-mechanical
fluctuations enhance repulsion between bimolecular lay-
ers," Proc. Natl. Acad. Sci. U. S. A. 83, 7132 (1986).
[18] J. N. Israelachvili and H. Wennerstroem, "Entropic forces
between amphiphilic surfaces in liquids," J. Phys. Chem.
96, 520–531 (1992).
[19] J. O. Radler, T. J. Feder, H. H. Strey, and E. Sack-
mann, "Fluctuation analysis of tension-controlled undu-
lation forces between giant vesicles and solid substrates,"
Phys. Rev. E 51, 4526–4536 (1995).
[20] U. Seifert, "Configurations of fluid membranes and vesi-
cles," Adv. Phys. 46, 13–137 (1997).
[21] B. G. Lorz, A.-S. Smith, C. Gege, and E. Sackmann,
"Adhesion of giant vesicles mediated by weak binding
of sialyl-lewisx to e-selectin in the presence of repelling
poly(ethylene glycol) molecules," Langmuir 23, 12293–
12300 (2007).
[22] S. A. Safran, N. Gov, A. Nicolas, U. S. Schwarz, and
T. Tlusty, "Physics of cell elasticity, shape and adhesion,"
Physica A 352, 171 (2005).
[23] T. Betz, M. Lenz, J.-F. Joanny, and C. Sykes, "Atp-
dependent mechanics of red blood cells," Proc. Natl.
Acad. Sci. U. S. A. 106, 15320–15325 (2009).
[24] C. Monzel, D. Schmidt, C. Kleusch, D. Kirchenbuchler,
U. Seifert, A.-S. Smith, K. Sengupta, and R. Merkel,
"Measuring
of bio-
membranes with dynamic optical displacement spec-
troscopy," Nat. Commun. 6, 8162 (2015).
stochastic displacements
fast
[25] J. Prost,
J.-B. Manneville,
and R. Bruinsma,
"Fluctuation-magnification of non-equilibrium mem-
branes near a wall," Eur. Phys. J. B 1, 465–480 (1998).
[26] S. Ramaswamy, J. Toner, and J.s Prost, "Nonequilib-
rium fluctuations, traveling waves, and instabilities in
active membranes," Phys. Rev. Lett. 84, 3494 (2000).
[27] J.-B. Manneville, P. Bassereau, S. Ramaswamy,
and
J. Prost, "Active membrane fluctuations studied by mi-
cropipet aspiration," Phys. Rev. E 64, 021908 (2001).
[28] P. Girard, J. Prost, and P. Bassereau, "Passive or active
fluctuations in membranes containing proteins," Phys.
Rev. Lett. 94, 088102 (2005).
[29] N. S. Gov and S. A. Safran, "Red blood cell mem-
brane fluctuations and shape controlled by atp-induced
cytoskeletal defects," Biophys. J. 88, 1859–1874 (2005).
[30] L. C.-L. Lin, N. Gov, and F. L. H. Brown, "Nonequilib-
rium membrane fluctuations driven by active proteins,"
J. Chem. Phys. 124, 074903 (2006).
[31] B. Loubet, U. Seifert, and M. A. Lomholt, "Effective ten-
sion and fluctuations in active membranes," Phys. Rev.
E 85, 031913 (2012).
[32] Y. Hanlumyuang, L. P. Liu, and P. Sharma, "Revisiting
the entropic force between fluctuating biological mem-
branes," J. Mech. Phys. Solids 63, 179–186 (2014).
[33] R. Alert, J. Casademunt, J. Brugu´es,
and P. Sens,
"Model for probing membrane-cortex adhesion by mi-
cropipette aspiration and fluctuation spectroscopy," Bio-
phys. J. 108, 1878–1886 (2015).
[34] D. Mizuno, C. Tardin, C. F. Schmidt, and F. C. MacK-
intosh, "Nonequilibrium mechanics of active cytoskeletal
networks," Science 315, 370–373 (2007).
[35] E. Ben-Isaac, Y. Park, G. Popescu, F. L. H. Brown,
N. S. Gov,
and Y. Shokef, "Effective temperature of
red-blood-cell membrane fluctuations," Phys. Rev. Lett.
106, 238103 (2011).
[36] H. Turlier, D. A. Fedosov, B. Audoly, T. Auth, N. S.
Gov, C. Sykes, J.-F. Joanny, G. Gompper, and T. Betz,
"Equilibrium physics breakdown reveals the active na-
ture of red blood cell flickering," Nat. Phys.
(2016),
10.1038/nphys3621.
8
[37] N. Gov and S. Safran, "Pinning of fluid membranes by
periodic harmonic potentials," Phys. Rev. E 69, 011101
(2004).
[38] L. C.-L. Lin and F. L. H. Brown, "Dynamic simulations
of membranes with cytoskeletal interactions," Phys. Rev.
E 72, 011910 (2005).
[39] E. Reister, T. Bihr, U. Seifert, and A.-S. Smith, "Two
intertwined facets of adherent membranes: membrane
roughness and correlations between ligand–receptors
bonds," New J. Phys. 13, 025003:1–15 (2011).
[40] J. Hu, R. Lipowsky, and T. R. Weikl, "Binding constants
of membrane-anchored receptors and ligands depend
strongly on the nanoscale roughness of membranes,"
Proc. Natl. Acad. Sci. U. S. A. 110, 15283–15288 (2013).
[41] T. Bihr, U. Seifert, and A.-S. Smith, "Multiscale ap-
proaches to protein-mediated interactions between mem-
branes-relating microscopic and macroscopic dynamics
in radially growing adhesions," New J. Phys. 17, 083016
(2015).
[43] J. A. Janes, H. Stumpf, Schmidt D., Seifert U.,
[42] R. Bruinsma, M. Goulian, and P. Pincus, "Self-assambly
of membrane junctions," Biophys. J. 67, 746–750 (1994).
and
Smith A.-S., "Statistical mechanics of an elastically
Static profile and correlations,"
pinned membrane:
ArXiv e-prints
(2018), arXiv:1806.05109 [physics.bio-
ph].
[44] L. C.-L. Lin and F. L. H. Brown, "Dynamics of pinned
membranes with application to protein diffusion on the
surface of red blood cells," Biophys. J. 86, 764–780
(2004).
[45] D. Schmidt, T. Bihr, U. Seifert, and A.-S. Smith, "Co-
existence of dilute and densely packed domains of ligand-
receptor bonds in membrane adhesion," Europhys. Lett.
99, 38003 (2012).
[46] M. Doi and S.F. Edwards, The Theory of Polymer Dy-
namics, Vol. 73 (Oxford University Press, 1988).
[47] R. Granek, "From semi-flexible polymers to membranes:
Anomalous diffusion and reptation," Journal de Physique
II 12, 1761 (1997).
[48] L. C.-L. Lin and F. L. H. Brown, "Simulating mem-
brane dynamics in nonhomogeneous hydrodynamic envi-
ronments," J. Chem. Theory Comput. 2, 472–483 (2006).
[49] E. Reister-Gottfried, K. Sengupta, B. Lorz, E. Sack-
mann, U. Seifert, and A.-S. Smith, "Dynamics of specific
vesicle-substrate adhesion: From local events to global
dynamics," Phys. Rev. Lett. 101, 208103:1–4 (2008).
[50] J. K. Sigurdsson, F. L. H. Brown, and P. J. Atzberger,
"Hybrid continuum-particle method for fluctuating lipid
bilayer membranes with diffusing protein inclusions,"
Journal of Computational Physics 252, 65 – 85 (2013).
[51] U. Seifert, "Dynamics of a bound membrane," Phys. Rev.
E 49, 3124–3127 (1994).
[52] N. Gov, A. G. Zilman,
and S. Safran, "Cytoskeleton
confinement and tension of red blood cell membranes,"
Phys. Rev. Lett. 90, 228101 (2003).
[53] L. Bourdieu J. Robert F. C. MacKintosh D. Chatenay
E. Helfer, S. Harlepp, "Microrheology of biopolymer-
membrane complexes." Phys. Rev. Lett. 85, 457 (2000).
[54] J. P´ecr´eaux, H.-G. Dobereiner, J. Prost, J.-F. Joanny,
and P. Bassereau, "Refined contour analysis of giant unil-
amellar vesicles," Eur. Phys. J. E 13, 277–290 (2004).
[55] J. B. Huppa, M. Axmann, M. A. Mortelmaier, B. F. Lille-
meier, E. W. Newell, M. Brameshuber, L. O. Klein, G. J.
Schutz, and M. M. Davis, "TCR–peptide–MHC inter-
actions in situ show accelerated kinetics and increased
affinity," Nature 463, 963 (2010).
[56] S. Fenz, T. Bihr, D. Schmidt, R. Merkel, U. Seifert,
K. Sengupta, and A.-S. Smith, "Membrane fluctuations
mediate lateral interactions between cadherin bonds,"
Nature Physics 13, 906–913 (2017).
[57] Tomas D. Perez, Masako Tamada, Michael P. Sheetz,
and W. James Nelson, "Immediate-early signaling in-
duced by e-cadherin engagement and adhesion," Journal
of Biological Chemistry 283, 5014–5022 (2008).
9
|
1505.07408 | 1 | 1505 | 2015-05-27T17:26:22 | Enhanced Stability of the Model Mini-protein in Amino Acid Ionic Liquids and Their Aqueous Solutions | [
"physics.bio-ph",
"physics.chem-ph",
"q-bio.BM"
] | Using molecular dynamics simulations, the structure of model mini-protein was thoroughly characterized in the imidazolium-based amino acid ionic liquids and their aqueous solutions. We report that the mini-protein is more stable when AAIL is added as a cosolvent. Complete substitution of water by organic cations and anions further results in hindered conformational flexibility of the mini-protein. This observation suggests that AAILs are able to defend proteins from thermally induced denaturation. We show by means of radial distributions that the mini-protein is efficiently solvated by both solvents due to agood mutual miscibility. However, amino acid based anions prevail in the first coordination sphere of the mini-protein. | physics.bio-ph | physics | Enhanced Stability of the Model Mini-protein in Amino Acid Ionic Liquids
and Their Aqueous Solutions
Guillaume Chevrot,1 Eudes Eterno Fileti,2 and Vitaly V. Chaban2
1)MEMPHYS - Center for Biomembrane Physics, Department of Physics, Chemistry and
Pharmacy, University of Southern Denmark, Campusvej 55, 5230 Odense, Denmark.
2) Instituto de Ciência e Tecnologia, Universidade Federal de São Paulo, 12231-280, São José
dos Campos, SP, Brazil
Abstract. Using molecular dynamics simulations, the structure of model mini-protein was
thoroughly characterized in the imidazolium-based amino acid ionic liquids and their aqueous
solutions. We report that the mini-protein is more stable when AAIL is added as a cosolvent.
Complete substitution of water by organic cations and anions further results in hindered
conformational flexibility of the mini-protein. This observation suggests that AAILs are able
to defend proteins from thermally induced denaturation. We show by means of radial
distributions that the mini-protein is efficiently solvated by both solvents due to agood mutual
miscibility. However, amino acid based anions prevail in the first coordination sphere of the
mini-protein.
Key words: solvation; protein; ionic liquid; amino acid; molecular dynamics simulation.
TOC Graphic
Introduction
Efficient extraction and preservation of proteins constitute important goals of modern
biotechnology.1-6 The extracted and preserved proteins must retain their structure and
function, since many kinds of analysis are available in vitro only. This often appears
impossible due to a high sensitivity of protein configurations to local intermolecular
interactions. The three-dimensional structure of protein is maintained by means of hydrogen
bonds and multiple hydrophobic interactions of slightly varying strength.3,6 Electrostatic
interactions involving polar amino acid residues and monovalent inorganic ions (chloride,
sodium, potassium, etc) are also of high importance to stabilize proteins and provide
acceptable level of hydration of the binding sites. Even tiny change in the environment may
alter the fine balance of these interactions causing denaturation of protein.7 Denaturation
ultimately leads to unfolding and inactivation, but can also result in specific unnatural
arrangements, whose function is either absent or far from the expected one. Many chronic
diseases are linked to incorrect protein folding. The effect of such changes is close to
mutations and may be fatal for the living organism.
An extensive evidence exists that solvation plays a critical role in stability and function
of most proteins.6-11 Accurate spatial information about proteins is obtained from routine
diffraction experiments. Consistent interpretation of the diffraction density maps requires
computer simulations with atomistic precision and adequate sampling of the phase space in
the vicinity of the protein. In turn, the presence of macromolecular solute induces significant
and permanent effects on microscopic structure and transport properties of the surrounded
solvent and dissolved ions. Specific regions inside many proteins exist, where water cannot
penetrate easily. The density of solvent in these regions is, consequently, different from the
bulk density at given temperature and pressure. This feature is important to maintain the
required configuration of an entire protein, as was found via atomistic molecular dynamics
simulations.
Small fluctuations of local temperature and pressure can result in protein denaturation,
since they perturb an accurate free energy balance. These same factors apply when a protein is
extracted from the
living organism and handled
in vitro. Chemical modification,
immobilization, additional stabilizing agents and genetic modification are major techniques to
allow protein storage outside their natural environment.3 The required temperature regime
must be thoroughly maintained to prevent thermally induced structure alterations, most of
which are strictly irreversible due to energy barriers.
Room-temperature ionic liquids (RTILs) comprise an organic cation and organic or
inorganic anion.12-15 In quite a few cases, these ionic combinations result in low melting
points (below 373 K), which nevertheless comes at a price of a significantly high shear
viscosity (ca. 100 cP). RTILs exhibit unique physical and chemical properties including
negligible vapor pressure, non-flammability, high thermal and chemical stability, high ionic
conductivity and interesting solvation behavior. RTILs are currently probed in versatile fields
of chemical engineering, such as chemical sensors, electrochemical power sources, solution
chemistry, gas capture, separation applications and catalysis.16-20 Certain ionic liquids are
environmentally friendly, without any clearly determined toxicity in relation to living cells.
There exists experimental evidence that grafting of ether groups to RTILs decreases toxic
effects. Successful applications of RTILs in biotechnology are known involving biological
catalysis and protein conservation. It was argued that RTILs can serve as highly efficient
participants in versatile biochemical processes, in addition to providing a well tunable
reaction media. Water cannot be completely removed from the reaction environment,
therefore researchers have to deal with biphasic systems: RTIL + water. According to recent
reports, aqueous solutions of RTILs foster protein activity and protein stability. Some RTILs
prevent or inhibit protein aggregation and improve refolding. An ability of ionic liquids to
exclude aggregating of proteins (for instance, during folding) is an extremely important
feature. Such a behavior implies strong specific interactions between proteins and RTILs,
which may be sometimes more favorable than those between proteins and water. Interactions,
observed in the protein solvated in RTIL and water, are not perfectly understood currently,
whereas mostly speculative interpretations are available. Molecular dynamics simulations are
able to provide significant assistance along these lines.
Following the conclusions of Herrman and coworkers,21 solvent properties of solutions
change from typical electrolyte solution-like behavior to molten salt-like behavior as a
function of RTIL concentration. Phase transition can be hypothesized in this case.
Bhattacharyya and coworkers22 used femtosecond up-conversion to study solvation dynamics
of a probe covalently attached via the lone cysteine group to a protein, in the presence of
methylimidazolium RTIL. The average solvation time decreased from 650 ps to 260 ps, as 1.5
RTIL was supplied. The authors ascribed this observation to protein unfolding, which made
the probe more exposed. Rama Devi and coworkers23 concluded that stability of proteins and
their functional groups can be drastically modulated by the addition of cosolvents. The
reported transfer free energies for a model protein from water to alkyl ammonium-based
RTILs suggested that all those RTILs act as stabilizers. Yoshimura and coworkers24 described
structural change in chicken egg white lysozyme in the highly concentrated aqueous 1-butyl-
3-methylimidazolium nitrate solutions (0-24M) by optical spectroscopy and small-angle X-
ray scattering methods. The protein adopted a partially globular state (tertiary structure was
disrupted), whereas the protein aggregation was inhibited. The partially globular state was
explained by dehydration of the important protein binding sites at high contents of RTIL.
Figueiredo and coworkers25 investigated protein destabilization incurred by 1-butyl-3-
methylimidazolium and 1-ethyl-3-methylimidazolium chloride and dicyanamide employing
molecular dynamics simulations, NMR spectroscopy and differential scanning calorimetry.
According to these authors, stabilization or destabilization of proteins depends on the
hydrophilicity of the RTIL anions. Binding of weakly hydrated anions to positively charged
or polar residues leads to partial dehydration of the backbone groups and is critical to control
protein stability. Thus, dicyanamide is more denaturating than chloride, which should not
depend on the cation and, to a reasonable extent, on the concentration.
Amino acid-based
ionic
liquids (AAILs) represent an
interesting and recent
implementation of RTILs.26-31 In AAILs, the anion is obtained from a molecular amino acid
via deprotonation, whereas the cation can be picked up in view of prospective usage. Being in
tight evolutional connection to alpha amino acids, AAILs are not toxic. They exhibit
relatively low ionic conductivities and high shear viscosities, which can be tuned via the
addition of water. Miscibility of water and AAILs is excellent over a wide range of contents.28
Dagade and coworkers29 reported thermodynamics of ionic hydration and discussed solvent-
solute interactions of AAILs in water at room conditions. Woo and coworkers31 simulated
certain equilibrium and transport properties of AAILs. We developed a simple and efficient
force field for a significant set of amino acid anions and 1-ethyl-3-methylimidazolium
cation.27
This work reports a model mini-protein solvation in the three pure AAILs (1-ethyl-3-
methylimidazolium alanine,
[EMIM][ALA]; 1-ethyl-3-methylimidazolium methionine,
[EMIM][MET]; 1-ethyl-3-methylimidazolium tryptophane, [EMIM][TRP]) and their aqueous
solutions. Classical empirical-potential molecular dynamics simulation offering an atomistic
precision of the results was employed as a primary research technique. The force field of
AAILs was developed recently by two of us.27 We describe structure of the mixed solvent in
the vicinity of mini-protein and characterize preferential solvation. We conclude that
solvation by AAILs is definitely favorable for the investigated mini-protein, since it better
preserves its genuine structure than water.
Methodology
Constant-temperature constant pressure molecular dynamics simulations of seven
principal systems (Table 1) were conducted at 310 K and 1 bar. Dynamics of the mini-protein
and surrounding solvent was recorded during 100 ns. The immediate coordinates of all
species were saved every 10 ps, whereas the immediate velocities and forces were not saved.
The equations-of-motion were propagated with a time-step of 0.002 ps. The time-step of
0.002 ps requires that the lengths of all intra-molecular (covalent) bonds involving hydrogen
atoms were fixed during dynamics. We applied the LINCS algorithm32 to achieve this. The
CHARMM36 force field33 was used to represent the mini-protein and the recently developed
force field was used to represent AAILs.27 Both force fields are fully compatible. One
chloride anion was added to the system to neutralize the mini-protein. Prior to productive
runs, the simulated systems were thoroughly equilibrated as described below.
Table 1. List of the simulated systems and their compositions
# system
AAIL
x (AAIL), mol%
# ion pairs
# water molecules
1
2
3
4
5
6
7
[EMIM][ALA]
[EMIM][ALA]
[EMIM][MET]
[EMIM][MET]
[EMIM][TRP]
[EMIM][TRP]
water
100
5
100
5
100
5
0
600
300
500
200
350
200
0
0
5700
0
0
3800
3800
5065
Equilibration of the macromolecule containing systems deserves an extensive attention,
since these systems possess lots of local free energy minima. These minima must be avoided
ultimately driving the system over a variety of possible states. With the above precautions in
mind, we performed the equilibration in the three stages. First, initial geometries were
prepared using PACKMOL.34 We used entry 1L2Y from the Protein Data Bank (PDB) for the
TRP-cage mini-protein (Figure 1). Molecular dynamics simulations of 100 ps with no
pressure coupling were performed on these geometries using GROMACS 5.0.2 with three-
dimensional periodic conditions. The geometry of the mini-protein was restrained at this stage
to defend it from any artificially induced conformational changes. Second, the barostat was
turned on and the simulations of 10 ns each were conducted at 400 K. The temperature of
simulation was temporarily increased to foster equilibration of the AAIL phase. Third, the
mini-protein atoms were relieved, and the entire system was simulated for 10 ns at 310 K and
1 bar.
Figure 1. The stick-and-ribbon superimposed representation of the TRP-cage protein
(sequence NLYIQ WLKDG GPSSG RPPPS).
The electrostatic interactions in all systems were simulated using direct Coulomb law
up to 1.2 nm and using the computationally efficient implementation of the Ewald method
(Particle-Mesh-Ewald) beyond 1.2 nm.35 The Lennard-Jones interactions were smoothly
brought down to zero from 1.0 to 1.2 nm using the switched potential approach. The constant
temperature was maintained using
the Bussi-Donadio-Parinello velocity
rescaling
thermostat36 with a relaxation time of 0.1 ps. The constant pressure was maintained using the
Parrinello-Rahman barostat37 with a relaxation time of 1 ps and the compressibility constant
of 4.5×10-5 bar-1. The molecular trajectories were analyzed by means of conventional
GROMACS utilities and the MDTraj software package38 using standard definitions of the
physical properties of interest. The images have been produced using VMD (Visual
Molecular Dynamics) software, version 1.9.39
Results and Discussion
The molecular configurations of
the
investigated
three-component
systems
corresponding to free energy minimum at 310 K and 1 bar are depicted in Figure 2. Water and
[EMIM][ALA] appear perfectly miscible. Our recent study indicates that AAILs exist as
mainly small ionic clusters in water, while water greatly decreases shear viscosity of AAILs.28
Good mutual miscibility of the simulated cosolvents likely means that both of them are
present in the first solvation shell of the mini-protein.
Figure 2. Final molecular configurations of the two selected simulated systems: (left) the
mini-protein in system 2 [x([EMIM][ALA]) = 5 mol%]; (right) the mini-protein in system 1
(pure [EMIM][ALA]). The 1-methyl-3-ethylimidazolium cations are blue, the alanine anions
are yellow, the water molecules are red. The mini-protein is in cyan and pink.
Figure 3 summarizes radial distribution functions (RDFs) computed between oxygen
atoms belonging to water and oxygen atoms belonging to the carboxyl group of the serine
residue. The serine residue was chosen because it is polar and, therefore, its strong interaction
with polar solvent is expected. Remember that both our cosolvents are polar. Indeed, the first
peak is located at 0.28 nm and also the second peak is distinguishable (although small by
height) at 0.48 nm. The height of the first peak is smaller in the case of pure water (system 7)
than in AAILs (systems 1-6). This trend constitutes a common observation in aqueous
solutions. Pure water possesses a very well structured network of hydrogen bonds, which is
responsible for high water-water RDF peaks. Interactions with the solute may be strong
themselves, but they are always weaker than water-water interactions. That is why local
density increase (RDF peak) in the vicinity of protein in pure water is modest. When a
significant concentration of another cosolvent is added, it perturbs hydrogen bonds of water,
especially upon creation of new hydrogen bonds with water. The resulting water structure in
the mixed solvent is weaker than that in pure water. Its binding to the protein produces a
higher peak even though the strength of the water-protein interaction remains intact.
The heights of the first peaks in the 5 mol% AAIL solutions are as follows:
[EMIM][TRP]>[EMIM][MET]>[EMIM][ALA]. This trend can be correlated to mobility of
cosolvents. [EMIM][TRP] is the least mobile AAIL due the anion size, [EMIM][MET] is
more mobile, and [EMIM][ALA] is most mobile due to the very small amino acid radical.
Coordination number of water in the first coordination sphere of serine amounts to 3.9
(system 7). Coordination numbers of the mini-protein with respect to the anions, n–=3.7 in all
AAILs (systems 2, 4, 6). One can conclude that AAIL substitutes most water molecules in the
vicinity of the mini-protein in the 5 mol% AAIL solutions.
Figure 3. Radial distribution functions (RDF) featuring spatial correlation between carboxyl
oxygen atoms of the serine residue (mini-protein) and oxygen atoms (water) in systems 2, 4,
and 6.
We selected an intrinsically acidic hydrogen atom of the imidazole ring to represent
binding behavior of the 1-ethyl-3-methylimidazolium cation. This site of the cation was
selected, because it is most electron deficient among atoms of [EMIM]+. Particularly this
atom participates in hydrogen bonding with other chemical entities including anions in the
condensed phase of ionic liquid. Figure 4 depicts RDFs between serine residue (carboxyl
oxygen atom) and imidazole hydrogen atom computed in the six simulated AAIL containing
systems. The amino acid anion of AAIL impacts significantly and provides different RDFs in
each AAIL.
Figure 4. Radial distribution functions (RDF) featuring spatial correlation between carboxyl
oxygen atoms of serine residue (mini-protein) and intrinsically acidic hydrogen (imidazole
ring) in systems 1, 2, 3, 4, 5, and 6.
The first peaks in all AAIL containing systems is located at 0.18 nm, which corresponds
to a strong hydrogen bond between the imidazolium-based cation and the mini-protein. The
first minima are poorly defined ranging from 0.3 to 0.4 nm. As composition of the first
coordination sphere directly depends on the definition of it, the corresponding coordination
numbers with respect to 1-ethyl-3-methylimidazolium cation, n+, must be treated with
caution. In the 5% AAIL solutions, n+=0.11 in [EMIM][ALA], n+=0.11 in [EMIM][MET] and
n+=0.07 in [EMIM][TRP]. Compare with n– reported above. The mutual affinity between the
mini-protein and amino acid anions are hereby proven.
In the pure AAIL systems (1, 3, 5), n+ obviously increases: n+ ([EMIM[ALA])=0.92,
n+ ([EMIM[ALA])=0.77, n+ ([EMIM[ALA])=0.23. This is in concordance with RDFs for the
mini-protein and the anions suggesting that affinity of the anions decreases in the following
row: [EMIM][TRP]>[EMIM][MET]>[EMIM][ALA].
The root-mean-square deviation (RMSD) of the macromolecule structure is a central
measure of its stability, in the broad sense, vs. time. Rupture of the initial mini-protein
structure would result in high values of RMSD. In turn, smaller RMSD values would suggest
a better conservation of the protein. RMSD is not only the function of solvent transport
properties, but also – and to a larger extent – on specific chemical interactions between the
mini-protein and solvents. According to Figure 5, the mini-protein has the largest
conformational flexibility in water. Water is a natural environment for proteins. Although the
investigated mini-protein retains stability at 310 K, temperature elevation will likely violate
an initial structure drastically. It is widely known that many proteins in the multicellular
organisms start denaturing at 40-50 °C. Such a denaturation is a primary reason that very little
higher animal are able to survive even short temperature elevations. Bacteria and virus are
much more advanced in this aspect, because their proteins are different.
Thermal denaturation is considered the most frequent denaturation pathway, which
takes place during storage of macromolecules in vitro. The conventional techniques, such as
chemical modification, immobilization, stabilizing agents, cannot defend from thermal
denaturation. However, this can be realistically achieved through the usage of different
solvent, because RMSD in pure AAILs is significantly smaller than in water and in 5 mol%
aqueous mixtures of AAILs. The difference between the cases of pure [EMIM][ALA],
[EMIM][MET], and [EMIM][TRP] is not drastic. All three AAILs efficiently solvate the
mini-protein similarly to water. They do not induce any observable structural alterations.
Moreover, thermal fluctuations of the mini-protein structure in AAILs are smaller suggesting
a better thermal stability of the mini-protein.
Figure 5. Evolution of the root-mean-square deviation (RMSD) of the mini-protein structure
upon the 100 ns long equilibrium molecular dynamics simulations. The first equilibrated
geometry was taken as a reference point. See legends for system designation.
Radius of gyration (Rg, Figure 6) confirms conclusions about a higher stability of the
mini-protein in AAILs and 5 mol% AAIL solutions in water. Little changes of Rg were
observed during 100 ns of equilibrium molecular dynamics simulations. Rg of the mini-
protein in [EMIM][ALA] equals to 0.74±0.01 nm; 0.76±0.01 nm in [EMIM][MET];
0.74±0.01 nm in [EMIM][TRP]; 0.74±0.01 nm in pure water. The standard deviation of Rg in
water is slightly larger than in other systems, which is consistent with larger RMSD
(Figure 5). All corresponding values of Rg and their components are summarized in Table S1.
Figure 6. Radius of gyration of the mini-protein as a function of time during equilibrium
molecular dynamics simulations. See legends for system designation. The average values of
radius of gyration are summarized in Table S1.
One of the most interesting questions that must be attended in this investigation is
which solvent and which chemical entity in particular prevail in the first coordination sphere
of the model mini-protein and are, therefore, responsible for its efficient solvation and
conservation. The posed question can be answered in terms of distribution functions and
coordination numbers, which are the integrals of those functions. Figures 7-8 investigate
spatial correlations between terminal hydrogen atoms (position w) and methyl group
hydrogen atoms of the mini-protein with oxygen atoms of water and three AAILs during the
100 ns long equilibrium molecular dynamics simulations of the seven systems. The two most
representative plots are shown here. An interested reader is referred to Supplementary
Information, Figures S1-S4, for all summarized RDFs. The w-H atoms are strongly correlated
with carboxyl groups of the anions by creating multiple hydrogen bonds with the length of
ca. 0.18 nm. In turn, both imidazolium cations and water molecules are pushed to the second
coordination sphere, where they exhibit quite smashed RDF peaks. Unlike [EMIM]+, certain
amount of water molecules is still present in the first shell, but the corresponding RDF
maximum is much smaller as compared to that of the anions.
5% [EMIM][TRP] AAIL/water mixture (right). Snapshots highlight mini-protein solvation.
Figure 7. Radial distribution functions (RDFs) featuring spatial correlation between hydrogen
atoms (position w) of the mini-protein and oxygen atoms of solvents in pure AAIL (left) and
Binding of all solvents to the methyl group is sufficiently different. Despite the first
peak corresponding to [EMIM]+ at 0.42 nm, no strong structure correlations were observed.
Even this peak vanishes in the 5 mol% [EMIM][MET] aqueous solution. The RDFs in other
AAILs (Figures S1-S2) completely support this conclusion. All applicable coordination
numbers range between 0.2 and 0.3. That is the methyl group appears insufficiently solvated
by both solvents. Based on the RDF in the pure [EMIM][TRP] system, one can identify that
310 K is lower than a glassy transition temperature for the system containing the model mini-
protein and this AAIL (system 5).
Figure 8. Radial distribution functions (RDFs) featuring spatial correlation between hydrogen
atoms (methyl groups) of the mini-protein and oxygen atoms of solvents in pure AAIL (left)
and 5% [EMIM][MET] AAIL/water mixture (right). Snapshots highlight mini-protein
solvation.
Conclusions
Classical molecular dynamics simulations using the recently developed force field for
amino acid-based ionic liquids were carried out to observe their solvation behavior in relation
to model mini-protein. In addition to pure AAILs, 5 mol% aqueous solutions of these AAILs
were considered, since water is always abundant in real biological systems. We found that
AAILs efficiently coordinate the mini-protein. This happens primarily thanks to amino acid
based anions, which interact strongly with (the ‘like dissolves like’ rule holds). AAILs push
some water molecules outside the first coordination sphere. This observation is in
concordance with previous reports.7,10 Being coordinated by the ions, the mini-protein
exhibits hindered mobility, as indicated by the RMSD analysis in all six AAIL containing
systems. The conformation flexibility is even smaller when water is completely excluded.
This observation favors applications of amino acid-based ionic liquids for conservation of
proteins and their subsequent usage in vitro. Importantly, solvation by AAILs does not
perturb an initial shape and configuration of the mini-protein, as suggested by constant radius
of gyration and its components. According to our computational analysis, AAILs emerge as
an interesting candidate solvent or cosolvent for a robust protein solvation, conservation and
storage. We anticipate that our paper will inspire more experimental efforts in this direction.
Acknowledgments
The research was supported by CAPES, FAPESP and CNPq.
Supporting Information
Table S1 summarizes radii of gyration and their components. Figures S1-S4 show radial
distribution functions for the mini-protein and selected AAILs.
Author Information
E-mails for correspondence: [email protected] (G.C.);
[email protected] (E.E.F.);
[email protected] (V.V.C.)
Zhang, T.; Wang, L. J.; Chen, H. Protein Extraction of the Long-Term Room
(1)
Temperature Storage Wheat Straw. Material Design, Processing and Applications, Parts 1-
4 2013, 690-693, 1252-1255.
REFERENCES
(2)
Fabian, C.; Ju, Y. H. A Review on Rice Bran Protein: Its Properties and Extraction
Methods. Critical Reviews in Food Science and Nutrition 2011, 51, 816-827.
Patel, R.; Kumari, M.; Khan, A. B. Recent Advances in the Applications of Ionic
(3)
Liquids in Protein Stability and Activity: A Review. Applied Biochemistry and
Biotechnology 2014, 172, 3701-3720.
Chen, X. W.; Mao, Q. X.; Wang, J. H. Ionic Liquids in Extraction/Separation of
(4)
Proteins. Progress in Chemistry 2013, 25, 661-668.
(5)
Asenjo, J. A.; Andrews, B. A. Aqueous Two-Phase Systems for Protein Separation:
A Perspective. Journal of Chromatography A 2011, 1218, 8826-8835.
Takekiyo, T.; Yamazaki, K.; Yamaguchi, E.; Abe, H.; Yoshimura, Y. High Ionic Liquid
(6)
Concentration-Induced Structural Change of Protein in Aqueous Solution: A Case Study
of Lysozyme. Journal of Physical Chemistry B 2012, 116, 11092-11097.
Vasantha, T.; Attri, P.; Venkatesu, P.; Devi, R. S. R. Structural Basis for the
(7)
Enhanced Stability of Protein Model Compounds and Peptide Backbone Unit in
Ammonium Ionic Liquids. Journal of Physical Chemistry B 2012, 116, 11968-11978.
Sun, Y.; Dominy, B. N.; Latour, R. A. Comparison of Solvation-Effect Methods for
(8)
the Simulation of Peptide Interactions with a Hydrophobic Surface. Journal of
Computational Chemistry 2007, 28, 1883-1892.
(9)
Sun, Y.; Latour, R. A. Comparison of Implicit Solvent Models for the Simulation of
Protein-Surface Interactions. Journal of Computational Chemistry 2006, 27, 1908-1922.
(10) Chowdhury, R.; Sen Mojumdar, S.; Chattoraj, S.; Bhattacharyya, K. Effect of Ionic
Liquid on the Native and Denatured State of a Protein Covalently Attached to a Probe:
Solvation Dynamics Study. Journal of Chemical Physics 2012, 137.
(11) Das, D. K.; Das, A. K.; Mandal, A. K.; Mondal, T.; Bhattacharyya, K. Effect of an Ionic
Liquid on the Unfolding of Human Serum Albumin: A Fluorescence Correlation
Spectroscopy Study. Chemphyschem 2012, 13, 1949-1955.
(12) Eilmes, A.; Kubisiak, P. Stability of Ion Triplets in Ionic Liquid/Lithium Salt
Solutions: Insights from Implicit and Explicit Solvent Models and Molecular Dynamics
Simulations. Journal of Computational Chemistry 2015, 36, 751-762.
(13) Bessac, F.; Maseras, F. Dft Modeling of Reactivity in an Ionic Liquid: How Many
Ion Pairs? Journal of Computational Chemistry 2008, 29, 892-899.
(14) Carlin, C.; Gordon, M. S. Ab Initio Calculation of Anion Proton Affinity and
Ionization Potential for Energetic Ionic Liquids. Journal of Computational Chemistry
2015, 36, 597-600.
(15) Addicoat, M. A.; Fukuoka, S.; Page, A. J.; Irle, S. Stochastic Structure Determination
for Conformationally Flexible Heterogenous Molecular Clusters: Application to Ionic
Liquids. Journal of Computational Chemistry 2013, 34, 2591-2600.
(16) Corvo, M. C.; Sardinha, J.; Menezes, S. C.; Einloft, S.; Seferin, M.; Dupont, J.;
Casimiro, T.; Cabrita, E. J. Solvation of Carbon Dioxide in [C(4)Mim][Bf4] and
[C(4)Mim][Pf6]
Ionic Liquids Revealed by High-Pressure Nmr Spectroscopy.
Angewandte Chemie-International Edition 2013, 52, 13024-13027.
(17) Chaban, V. Hydrogen Fluoride Capture by Imidazolium Acetate Ionic Liquid.
Chemical Physics Letters 2015, 625, 110-115.
(18) Ohba, T.; Chaban, V. V. A Highly Viscous Imidazolium Ionic Liquid inside Carbon
Nanotubes. Journal of Physical Chemistry B 2014, 118, 6234-6240.
(19) Hantal, G.; Voroshylova, I.; Cordeiro, M. N. D. S.; Jorge, M. A Systematic Molecular
Simulation Study of Ionic Liquid Surfaces Using Intrinsic Analysis Methods. Physical
Chemistry Chemical Physics 2012, 14, 5200-5213.
(20) Chaban, V. V.; Prezhdo, O. V. Water Phase Diagram Is Significantly Altered by
Imidazolium Ionic Liquid. Journal of Physical Chemistry Letters 2014, 5, 1623-1627.
(21) Weingartner, H.; Cabrele, C.; Herrmann, C. How Ionic Liquids Can Help to Stabilize
Native Proteins. Phys Chem Chem Phys 2012, 14, 415-426.
(22) Chowdhury, R.; Sen Mojumdar, S.; Chattoraj, S.; Bhattacharyya, K. Effect of Ionic
Liquid on the Native and Denatured State of a Protein Covalently Attached to a Probe:
Solvation Dynamics Study. J Chem Phys 2012, 137, 055104.
(23) Vasantha, T.; Attri, P.; Venkatesu, P.; Devi, R. S. Structural Basis for the Enhanced
Stability of Protein Model Compounds and Peptide Backbone Unit in Ammonium Ionic
Liquids. J Phys Chem B 2012, 116, 11968-11978.
(24) Takekiyo, T.; Yamazaki, K.; Yamaguchi, E.; Abe, H.; Yoshimura, Y. High Ionic Liquid
Concentration-Induced Structural Change of Protein in Aqueous Solution: A Case Study
of Lysozyme. J Phys Chem B 2012, 116, 11092-11097.
(25) Figueiredo, A. M.; Sardinha, J.; Moore, G. R.; Cabrita, E. J. Protein Destabilisation in
Ionic Liquids: The Role of Preferential Interactions in Denaturation. Phys Chem Chem
Phys 2013, 15, 19632-19643.
(26) Ohno, H.; Fukumoto, K. Amino Acid Ionic Liquids. Accounts of Chemical Research
2007, 40, 1122-1129.
(27) Fileti, E. E.; Chaban, V. V. The Scaled-Charge Additive Force Field for Amino Acid
Based Ionic Liquids. Chemical Physics Letters 2014, 616, 205-211.
(28) Chaban, V. V.; Fileti, E. E. Ionic Clusters Vs Shear Viscosity in Aqueous Amino Acid
Ionic Liquids. J Phys Chem B 2015, 119, 3824-3828.
(29) Dagade, D. H.; Madkar, K. B.; Shinde, S. P.; Barge, S. S. Thermodynamic Studies of
Ionic Hydration and Interactions for Amino Acid Ionic Liquids in Aqueous Solutions at
298.15 K (Vol 117, Pg 1031, 2013). Journal of Physical Chemistry B 2013, 117, 9584-
9584.
(30) Zhu, J. F.; He, L.; Zhang, L.; Huang, M.; Tao, G. H. Experimental and Theoretical
Enthalpies of Formation of Glycine-Based Sulfate/Bisulfate Amino Acid Ionic Liquids.
Journal of Physical Chemistry B 2012, 116, 113-119.
(31) Sirjoosingh, A.; Alavi, S.; Woo, T. K. Molecular Dynamics Simulations of
Equilibrium and Transport Properties of Amino Acid-Based Room Temperature Ionic
Liquids. Journal of Physical Chemistry B 2009, 113, 8103-8113.
(32) Hess, B.; Bekker, H.; Berendsen, H. J. C.; Fraaije, J. G. E. M. Lincs: A Linear
Constraint Solver for Molecular Simulations. J. Comp. Chem. 1997, 18, 1463.
(33) Best, R. B.; Zhu, X.; Shim, J.; Lopes, P. E.; Mittal, J.; Feig, M.; Mackerell, A. D., Jr.
Optimization of the Additive Charmm All-Atom Protein Force Field Targeting Improved
Sampling of the Backbone Phi, Psi and Side-Chain Chi(1) and Chi(2) Dihedral Angles. J
Chem Theory Comput 2012, 8, 3257-3273.
(34) Martinez, L.; Andrade, R.; Birgin, E. G.; Martinez, J. M. Packmol: A Package for
Building Initial Configurations for Molecular Dynamics Simulations. J Comput Chem
2009, 30, 2157-2164.
(35) Darden, T.; York, D.; Pedersen, L. Particle Mesh Ewald: An N·Log(N) Method for
Ewald Sums in Large Systems. J. Chem. Phys. 1993, 98, 10089-10099.
(36) Bussi, G.; Donadio, D.; Parrinello, M. Canonical Sampling through Velocity
Rescaling. J. Chem. Phys. 126 2007, 126, 014101-014108.
(37) Parrinello, M.; Rahman, A. Polymorphic Transitions in Single Crystals: A New
Molecular Dynamics Method. J. Appl. Phys. 1981, 52, 7182-7192.
(38) McGibbon, R. T.; Beauchamp, K. A.; Schwantes, C. R.; Wang, L. P.; Hernandez, C. X.;
Harrigan, M. P.; Lane, T. J.; Swails, J. M.; Pande, V. S. Mdtraj: A Modern, Open Library for
the Analysis of Molecular Dynamics Trajectories. BioRxiv 2014.
(39) Humphrey, W.; Dalke, A.; Schulten, K. Vmd: Visual Molecular Dynamics. Journal of
Molecular Graphics 1996, 14, 33-38.
|
1905.03575 | 1 | 1905 | 2019-05-09T12:39:03 | A TDDFT-based Study on the Proton-DNA Collision | [
"physics.bio-ph",
"physics.chem-ph"
] | The interaction of heavy charged particles with DNA is of interest for several areas, from hadrontherapy to aero-space industry. In this paper, a TD-DFT study on the interaction of a 4 keV proton with an isolated DNA base pair was carried out. Ehrenfest dynamics was used to study the evolution of the system during and after the proton impact up to about 193 fs. This time was long enough to observe the dissociation of the target, which occurs between 80-100 fs. The effect of base pair linking to the DNA double helix was emulated by fixing the four O3' atoms responsible for the attachment. The base pair tends to dissociate into its main components, namely the phosphate groups, sugars and nitrogenous bases. A central impact with energy transfer of 17.9 eV only produces base damage while keeping the backbone intact. An impact on a phosphate group with energy transfer of about 60 eV leads to backbone break at that site together with base damage, while the opposite backbone site integrity is kept is this situation. As the whole system is perturbed during such a collision, no atom remains passive. These results suggest that base damage accompanies all backbone breaks since hydrogen bonds that keep bases together are much weaker that those between the other components of the DNA. | physics.bio-ph | physics | A TDDFT-based Study on the Proton-DNA
Collision
Rodrigo Seraide,† Mario A. Bernal,∗,† Gustavo Brunetto,† Umberto de
Giovannini,‡,¶ and Angel Rubio‡,§
†Instituto de F´ısica Gleb Wataghin. Universidade Estadual de Campinas. SP. Brazil.
‡Max Planck Institute for the Structure and Dynamics of Matter and Center for
Free-Electron Laser Science & Department of Physics. Luruper Chaussee 149. 22761
¶Dipartimento di Fisica e Chimica. Universit`a degli Studi di Palermo. Via Archirafi 36.
Hamburg. Germany.
§Nano-Bio Spectroscopy Group and ETSF, Dpto. F´ısica de Materiales, Universidad del
I-90123. Palermo. Italy.
Pa´ıs Vasco UPV/EHU, 20018 San Sebasti´an, Spain.
E-mail: [email protected]
9
1
0
2
y
a
M
9
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
5
7
5
3
0
.
5
0
9
1
:
v
i
X
r
a
1
Abstract
The interaction of heavy charged particles with DNA is of interest for several areas,
from hadrontherapy to aero-space industry. In this paper, a TD-DFT study on the
interaction of a 4 keV proton with an isolated DNA base pair was carried out. Ehrenfest
dynamics was used to study the evolution of the system during and after the proton
impact up to about 193 fs. This time was long enough to observe the dissociation of
the target, which occurs between 80-100 fs. The effect of base pair linking to the DNA
double helix was emulated by fixing the four O3' atoms responsible for the attachment.
The base pair tends to dissociate into its main components, namely the phosphate
groups, sugars and nitrogenous bases. A central impact with energy transfer of 17.9
eV only produces base damage while keeping the backbone intact. An impact on a
phosphate group with energy transfer of about 60 eV leads to backbone break at that
site together with base damage, while the opposite backbone site integrity is kept is this
situation. As the whole system is perturbed during such a collision, no atom remains
passive. These results suggest that base damage accompanies all backbone breaks since
hydrogen bonds that keep bases together are much weaker that those between the other
components of the DNA.
Introduction
The interaction of ionizing particles with DNA is a very complex process which depends
both on the particle track structure (radiation quality) and the genetic material geometrical
conformation. The early physicochemical damage that ionizing radiation induces in DNA
may lead to biological effects. These effects are of supreme importance for medical radiation
applications, both for diagnostic and therapeutic procedures. In addition, aerospace industry
is interested on this problem as astronauts are exposed to charged particle radiation during
their missions, and this includes heavy particles with mass even larger than a proton. The
radiobiological problem consists in studying biological effects induced by ionizing particles in
2
living beings. Several approaches have been used to deal with this problem during the past
seven decades. In vitro assays, in which cellular cultures are irradiated and later analyzed,
is the main source of information for understanding this problem. This is the case of the
pioneer works of Karl Sax and coworkers, 1 which were used by Lea and Catchside 2 as an
empirical base to formulate a successful biophysical model for the early DNA damage. Later,
Kellerer and Rossi proposed the long-standing Dual Radiation Action Theory, 3 which states
that lethal lesions induced by ionizing radiation in cells are produced by the interaction of
two sub-lesions (probably double strand breaks, DSB).
With the rapid increase of computing power along the last few decades, numerical ap-
proaches came out. For instance, Monte Carlo simulation of particle transport can be com-
bined with a DNA geometrical models and biophysical model in such a way that DNA
damage probability can be estimated. 4 -- 7 The latter approach counts a DNA damage, typical
a single strand break (SSB), when an energy deposition above a certain threshold value oc-
curs inside the target in question. Commonly, this target is the sugar-phosphate group. This
method implicitly assumes that the collision of the ionizing particle with DNA is a one-body
problem. That is, the rest of the DNA molecule remains frozen when the incoming particle
interact with the atom in question. This is not the case in reality, mainly when dealing with
relatively slow ions which produce a strong perturbation of the target system.
Time Dependent Density Functional Theory (TD-DFT) emerges as a powerful tool to
study the full dynamics of collisions involving complex systems since it is capable to ac-
count for the many body problem in a consistent way. First of all, the ground state of
the target system is determined using the Density Functional Theory (DFT). According to
the Hohenberg-Kohn theorem, the ground-state density is enough to determine the ground
state of an electronic system. 8 Kohn and Sham 9 found a way to uncouple the Schrodinger
equation system of the electronic system, making the problem easier to solve. That is, the
Kohn-Sham formalism is able to exactly map the interacting system into a non-interacting
and easier to solve problem.
3
In principle, TD-DFT can be used to study the collision between a charged particle and
DNA or some of its constituents. Bacchus et al. 10 studied the collision of carbon ions on
nitrogenous bases thymine, uracil and 5-halouracil. Targets were bombarded at various in-
cidence direction and impact parameters. Calculations were carried out with the MOLPRO
package. 11 They determined charge transfer cross sections for different carbon ions charge
states by following an impact parameter approximation. The authors speculate about dis-
sociation cross sections but they did not study this process directly. Sadr-Arani et al. 12 -- 15
have carried out several works using experimental and theoretical methods for studying the
fragmentation of DNA/RNA bases such as uracil, cytosine, adenine, and guanine. Their cal-
culations were based on the DFT formalism by they did not explicitly simulate any collision
process. Instead, they stretched bonds up to break and determined the involved dissocia-
tion energies and possible fragments. Lopez-Tarifa et al. 16 have recently used the TD-DFT
approach to study the fragmentation of doubly ionized uracil in gas phase. They did not
account for the explicit incidence of any projectile. They simply removed electrons from
inner-shells ad hoc and let the excited molecule evolve in time.
Classical molecular dynamics has been also used to study the collision of charged particles
with DNA. Albofath et al. 17 used the reactive force field ReaxFF 18 to study the role of
hydroxyl free radicals on DNA damage. They randomly distributed free hydroxyl radicals in
small pockets around a DNA fragment and followed the evolution of the system. They found
that OH radicals produce holes in the sugar-moiety rings and that they evolve to larger holes
that comprise several bases. Then this damage propagates to the bases and lead to single and
double strand breaks. One year later, Abolfath et al. 19 continued studying the same process
using the GEANT4-DNA Monte Carlo package 20 to obtain the initial position of hydroxyl
radicals. Then, the interaction of those radicals with DNA was described by the REAXFF-
based molecular dynamics approach. Primary 1 MeV electrons and protons were studied in
this work. They reported that protons produce four times more DNA double strand breaks
that electrons with the same energy. Bottlander et al. 21 used the REAX force field provided
4
by Abolfath et al. 19 to study the interaction of protons with DNA in a NaCl aqueous solution.
They simulated the direct interaction of a proton with a DNA fiber fragment by uniformly
distributing the energy transferred by the projectile to the target atoms within a cylinder
with 2 A radius. This energy was determined from the particle stopping power. The authors
reported the number of SSB and DSB produced by different energy transferred to the medium
when the projectile travels along the three main cartesian axes. The effect of a violent sock
wave created by ions with a very high stopping power (or linear energy transfer) has been
also studied using classical molecular dynamics. 22,23 They used the CHARMM potential
model to simulate the evolution of DNA atoms after the passage of the ion so explicit bond
breakage was not accounted for.
Instead, they estimated energy changes in DNA bonds
due the influence of the ion-induced shock wave and speculated on the possible creation of
single strand breaks. Recently, Bacchus-Montabonel and Calvo 24,25 studied the effect of the
hydration shell around biomolecular targets (uracil and aminooxazole) on the proton-induced
charge transfer process. This effect was done by adding only two water molecules at different
molecular sites. They determined charge transfer cross sections during the impact of 10 eV
to 10 keV protons using a software package based on the impact parameter approximation,
rather than using TD-DFT calculations.
This work aims at the study of the proton-DNA collision problem using the TD-DFT
to see how a base-pair evolves during and after the impact of an energetic proton. This
approach should allow the observation of many-body effects during this collision. In addition,
in-vacuum dissociation times and the energy required for this dissociation can be estimated
under different conditions, including different impact parameters and bounding with neighbor
base-pairs. It should be remarked that the detailed study of the DNA dissociation is out of
the scope of this work. We simply want to have a qualitative picture of this process as a
support for the introduction of a new approach to study the early DNA damage induced by
ionizing radiation. This new method would be an alternative to current biophysical models
(discussed above). That is, those approaches based on the assumption that only the atom
5
targeted by the incoming particle is affected while the others remain frozen and that double
strand breaks can be induced after the production of two close enough single strand breaks.
Up to our knowledge, this is the first time the TD-DFT approach is used to explicitly study
the collision between a heavy charged particle and a DNA base pair.
Atomic units are used throughout this work, unless otherwise stated.
Methods
Theoretical background of the TD-DFT.
The electronic Schrodinger equation of the interacting system with N electrons with positions
at (r1, r2, ..., rN), respectively, is
(cid:110) − 1
2
N(cid:88)
i=1
∇2
i +
1
2
2(cid:88)
i,j=1
N(cid:88)
i=1
vext(ri)
1
ri − rj +
(cid:111)
Ψ(r1, r2, ..., rN) = EΨ(r1, r2, ..., rN),
(1)
where the first term is the kinetic energy of electrons and the second one is the so-called
Hartree term. The external potential in absence of electromagnetic fields is
vext(ri) = − M(cid:88)
k=1
Zk
ri − RK
(2)
and comes from the interaction of electrons with point-like nuclei. After solving equation
(1), the electronic density n(r) can be determined as
(cid:90)
n(r) = N
d3r2 ··· d3rN Ψ(r, r2, ..., rN)Ψ∗(r, r2, ..., rN).
(3)
Equation (1) is very hard to solve but Kohn and Sham found a simpler and exact way
to solve it by introducing the exchange-correlation potential vxc(r). According to their
approach, the equation system (1) can be decomposed in equations for the orbitals φi(r)
6
forming a single Slater determinant of a fictitious non-interacting system with the same
density of the interacting one as following
∇2 + vHartree[n](r) + vext(r) + vxc[n](r)
φKS
i
(r) = EφKS
i
(r),
(4)
(cid:110) − 1
2
where
(cid:90)
d3r(cid:48) n(r(cid:48))
r − r(cid:48)
vHartree[n](r) =
(5)
and vxc[n](r) is the exchange-correlation potential which accounts for the many-body
effects of the problem. Notice that the Hartree and exchange-correlation potentials are
(cid:111)
(cid:111)
functional of the density defined as
N(cid:88)
i=1
n(r) =
φKS
i
(r)2.
For the time-dependent case, Kohn-Sham equations are
(cid:110) − 1
2
∇2 + vHartree[n](r, t) + vext(r, t) + vxc[n](r, t)
φKS
i
(6)
(r, t) = i
∂
∂t
φKS
i
(r, t),
(7)
Now both the density n(r, t) and nuclei positions RK(r, t) are functions of time. Similar
to the ground state calculation, equation system (7) is solved in a self-consistent way for each
time step. Here, we used the Adiabatic Local Density Approximation for describing the time-
dependent exchange-correlation functional vxc[n](r, t). 26 The time-evolution of nuclei was
described through the Ehrenfest dynamics. In this formalism, nuclei are treated classically
and allowed to move under the influence of the mean field generated by electrons. That is,
their equation of motion is
mK
∂2RK
∂t2 = −∇KV ( ¯R),
(8)
where mK and Rk are the nucleus mass and position, respectively. V ( ¯R) accounts for
7
the electron-nucleus attraction and nucleus-nucleus repulsion and is a function of the nuclei
positions ¯R = (R1, R2, ..., RM ). In this approximation the solution to the time dependent
Schrodinger equation is obtained propagating the classical equation of motion for the ions
(8) together with the quantum mechanical TDDFT equations for the electrons (7) until a
given time.
Collision setup
A proton with about 4 keV energy impacts on an isolated Guanine-Cytosine B-DNA base-
pair (bp) at rest. Atoms positions correspond to canonical B-DNA as that found in Protein
Data Bank ID 309D. 27 This bp contains the whole phosphate group on one side, ending
in O3', and the O3' atom belonging to the phosphate adjacent group.
In other words,
this bp's backbone ends in two O3' atoms. These terminal oxygen atoms were fixed in
some calculations to simulate the effect of bounding to the adjacent base pairs (see details
below). In order to stabilize the molecule, we completed the dangling bonds with hydrogens.
This means that hydrogen atoms were added to both O3' terminals and another one was
attached to the O2P atom, which is responsible for some DNA-protein binding. Two impact
parameters were included in this study. First, the proton impinges the DNA bp with 0 impact
parameter with respect to the molecule's geometrical center, near the hydrogen-bridge bonds
that link nitrogenous bases. Second, the proton impacts the bp with 0 impact parameter
with respect to the upper phosphorus atom, which belongs a the sugar-phosphate group.
The proton initially travels along the Z axis, which normally crosses the plane containing
the nitrogenous bases atoms. Fig (1) shows the localization of the DNA atoms and the
incoming proton. It should be remarked that this proton is treated as any other ion of the
target system, as described in eq. 8.
8
Figure 1: Proton-DNA collision setups.
Ground state calculation
The Octopus code version 4.1.2 28 -- 30 was used to carry out TDDFT calculations. A ground
state calculation for the DNA bp was done in a first stage. The Local Density Approximation
(LDA) and the modified Perdew& Zunger LDA 31 were used for the exchange and correlation
functionals, respectively. The system in question was placed inside a 46x20x20 a.u.3 box and
the calculation grid was obtained using a 0.4 a.u. spacing along the three Cartesian axes.
This spacing was found after an optimization process during which the system total energy
converged. Troullier-Martins pseudopotentials were used for all the atoms that conform the
bp 32 in such a way that their K-shell electrons were not treated explicitly. Using these
pseudopotentials, the number of orbitals for the whole bp was 119.
9
a)b)Guanine(GUA)Cytosine(CYT)sugar-phosphatebackbonesugar-phosphatebackboneH+H+CNOPHAtom TypesxzyTime-dependent calculations
After having obtained the ground state of the DNA bp, the proton was placed ad hoc
according to the impact parameter in question. For the proton with zero impact parameter
with respect to the geometrical center of the bp, the initial position was (0,0,-15) a.u. For
the proton with zero impact parameter w.r.t. the upper phosphorus atom (Cytosine side),
the initial point was (16.721,-4.396,-15) a.u. The initial proton velocity was (0,0,0.4) a.u. in
both cases so the proton impinges normal to the plane defined by the nitrogenous bases. A
first calculation was done for the central impact case in which all atoms were free to move.
In a second stage, the four O3' terminal atoms were fixed. This emulates the case in which
the bp is bound to a double helix DNA chain. Time step was set to 0.05 a.u. and the total
calculation time was 8000 a.u. (∼193 fs). This time was chosen in such a way that the initial
dissociation process can be observed. Calculations were carried out in a 120-core cluster and
a single 193 fs calculation took about 6 weeks of wall-clock time. Absorbing boundary
conditions were employed to prevent artificial reflections of electrons at the boundary of
the simulation box. 33 The complex potential method was used. According to preliminary
calculations, a temperature of 300 K shows negligible effects on the evolution of the bond
lengths in question. Time-evolution rendering was done with the VMD software. 34
Results
Figure 2 shows snapshots at characteristic times during the evolution of the DNA bp after
the proton impact at zero impact parameter where all atoms are free to move. Blue shaded
area represents the 0.001 % electronic density isosurface to show the evolution during the
collision. Snapshot a) captures the proton just passing through the bp. Snapshot b) displays
the proton charge capture. In the following lines, an energetic analysis will be carried out
as a consistency test of our calculations. Yet, it does not aim at a rigorous explanation
of the DNA dissociation. The energy transferred by the proton to the bp was 17.9 eV in
10
this collision. This energy is mainly transferred to the electrons during the collision and
about 33 % (6.02 eV) of it is subsequently transferred to the ions until just before the
dissociation process start. At first glance, it seems that the bp tends to dissociate into
its main components, namely the phosphate groups, sugars, and bases. However, Fig. 3a
shows that both O5' atoms dissociate from the corresponding phosphate groups. In fact,
they remain attached to the corresponding sugar through the C5' atom (ester bond), which
means that the integrity of the deoxyribose sugar prevails over that of the phosphate group.
P-O5' (P-O(C)) and P-O2 (P=O) bonds require about 3.67 eV and 6.03 eV for breaking,
respectively. 35 Thus it is more energetically advantageous to break the P-O5' bond instead of
the P-O2 one. It was also obtained that the P-O5' bond length is ∼1.593 A, which remains
stable until the dissociation process takes place. This value agrees with results reported
in the literature (1.591-1.603 A). 35 Then, about 7.34 eV would be used to break both P-
O5' bonds. The hydrogen-bonds that keep bases together are relatively weak. The binding
energy of the N-H-N and N-H-O bonds are ∼0.135 eV and ∼0.301 eV, respectively. 36 In the
CG bp, there are one N-H-N and two N-H-O bonds so the total binding energy for these
hydrogen-bonds is around 0.737 eV. Thus, an energy of about 8.077 eV is used to dissociate
the three hydrogen bonds and two sugar-phosphate bonds so still remain additional 9.823 eV
from the transferred energy, which include the kinetic energy transferred to the ions (6.02
eV). Figure 3a shows that the sugar-cytosine bond is broken unlike that between the sugar
and the guanine base, which remains stable during the calculation time. C-N bond energy
is about 3.158 eV 37 so it is estimated from these results that about 11.235 eV are used to
dissociate the DNA bp. The remaining transferred energy should be converted into kinetic
energy of the dissociation fragments.
Fig. 3a shows that the DNA bp actually dissociates into five products. Bond breaking
occurs so that the main constituents of the DNA separate from each others. That is, the
molecule tends to produce fragments such as phosphite groups, nitrogenous bases and de-
oxyriboses. This fragmentation pathway seems to be plausible since the fragments produced
11
are relatively stable radicals and the linkage between them should be the weakest bonds of
the molecule. However, unlike the sugar-cytosine bond, the sugar-guanine bond is stable
until 193 fs. Sugar-cytosine and cytosine-guanine bonds begin to dissociate almost simul-
taneously (from ∼50 fs on), while sugar-phosphate bonds take a bit longer (from ∼80 fs).
These dissociation times are consistent with those reported for large molecules (∼0.1 ps). 38
The time elapsed between the proton impact and the beginning of dissociation was estimated
as ∼49 fs. These results shows that passage of the proton perturbs the whole system. That
is, this is a many-body collision in which no component of the system remains frozen.
Figure 2: System snapshots for some important stages during the proton-free DNA central
collision: a) the proton just passing through the bp, b) the proton leaves the bp taking
a fraction of the system charge (electron capture process), c) beginning of the dissociation
process, and d) the bp dissociates with into five fragments. Animation with the whole process
can be found in supplementary information.
12
a)b)c)d)xzyGUAGUAGUAGUACYTCYTCYTCYT1.21 fs2.42 fs44.7 fs193 fsFigure 4 shows the same results as in Fig. 2, but with the O3' terminal atoms held fixed
in order to emulate the binding of the bp to their neighbors. The energy transfer is 17.9
eV again which means that the binding of the bp to their neighbors does not influence this
quantity, at least for central impact. This energy transfer occurs in a very short time so it
is possible that the effect of bp linking through oxygen atoms relatively far from the impact
region is very weak. As in the free DNA case, about 6.01 eV of this energy is later transferred
to the ions after the collision. The dissociation process can be better observed in Fig. 3b.
The dissociation of the hydrogen bonds begins almost at the same time as in the free bp
case but now the process seems to be slower. At the maximum calculation time, hydrogen
bonds lengths are larger for the bound bp than for the free one. Sugar-cytosine bond seems
to be in dissociation route but at a slower velocity, beginning at ∼100 fs and thus delayed
compared with the free configuration. Unlike the free bp situation, phosphate-sugar bonds
do not dissociate. This means that only the so-called base damage occurs while the DNA
backbone is not broken. The cytosine base is ejected while guanine remains attached to the
corresponding sugar. That is, only three fragments are produced in this case. This behavior
would be expected as the bp is now linked to their neighbors and the impact was on the
hydrogen bonds that keep bases together.
Finally, Fig. 5 shows snapshots of the proton DNA collision when the projectile impinges
the phosphorus atom located on the cytosine side. This is a head-on collision where the
proton transfers 61.8 eV to the bp. Unlike the two previous configurations, this is violent
impact against the phosphorus atom so that 30.74 eV are immediately transferred to this
ion, almost 50 % of the total energy transfer. This is an energy transfer high enough to even
break the P-O and P=O bonds, which requires about 3.67 eV and 6.03 eV for breaking,
respectively. At ∼2 fs and ∼10 fs, O1P and O2P atoms are ejected from the impacted
phosphate group. At ∼100 fs, even the phosphorus atom is emitted, together with another
oxygen atom and a proton. According to Fig. 3c, hydrogen bonds are dissociated, despite
that the impact is relatively far from this region. Only the O5' atom remains linked to the
13
sugar in this sugar-phosphate group. In addition, the sugar is dissociated from the cytosine
base, yet the process is slower than hydrogen-bonds dissociation. Again, the sugar-guanine
bond survive to the proton impact. C5'-C4' and C5-O5' in both sugars oscillates but do not
break so the deoxyribose integrity seems to be preserved.
(a)
(b)
(c)
Figure 3: Length of some important DNA bonds as a function of time during and after a
collision with a 4 keV proton for cases in which a) all atoms are free and the proton impacts
on the bp center, b) all atoms are free except four oxygen atoms that would link the bp to
their neighbors and the proton impacts on the bp center, and all atoms are free except four
oxygen atoms that would link the bp to their neighbors and c) the proton impacts on one
of the phosphorus atoms.
Discussion
According to this TD-DFT picture, DNA is initially ionized by the proton, both by electron
capture and direct ionization. Simultaneously, the proton perturbs the whole system (see
14
024681012140.0010.010.11101001000Phosphate-sugarBondlength(a.u.)time(fs)P-O5`,CytosineP-O5`,Guanine024681012Sugar-CytosineBondlength(a.u.)C1`-N10510152025Cytosine-GuanineBondlength(a.u.)N4-O6N3-N1O2-N2024681012Sugar-GuanineFreeDNAbase-pair.Centralimpact.Bondlength(a.u.)N9-C1`024681012140.0010.010.11101001000Phosphate-sugarBondlength(a.u.)time(fs)C5`-C4`,CytosineC5`-O5`,CytosineC5`-O5`,GuanineC5`-C4`,Guanine024681012Sugar-CytosineBondlength(a.u.)C1`-N10510152025Cytosine-GuanineBondlength(a.u.)N4-O6N3-N1O2-N2024681012Sugar-GuanineBoundDNAbase-pair.Centralimpact.Bondlength(a.u.)N9-C1` 0 2 4 6 8 10 12 14 0.001 0.01 0.1 1 10 100 1000Phosphate-sugar time (fs)C5`-C4`, CytosineC5`-O5`, CytosineP-O2P, CytosineP-O1P, CytosineC5`-O5`, GuanineC5`-C4`, Guanine 0 2 4 6 8 10 12Sugar-Cytosine (a.u.)C1`-N1 0 5 10 15 20 25Cytosine-Guanine (a.u.)N4-O6N3-N1O2-N2 0 2 4 6 8 10 12Sugar-GuanineBound DNA base-pair. Impact on upper P atom.Bond (a.u.)N9-C1`Figure 4: System snapshots for some important stages during the proton-bound DNA central
collision: a) the proton just passing through the bp, b) the proton leaves the bp taking a
fraction of the system charge (electron capture process), c) beginning of the dissociation
process, and d) the bp dissociates with into three fragments. Animation with the whole
process can be found in supplementary information.
supplementary information with animations) and this perturbation together with the system
ionization lead to the target molecule dissociation. The linking of the DNA bp to their
neighbor tends to keep the integrity of the DNA backbone when the proton impacts on the
bp center so only bases are damaged. Base damage was present in all the situation studied
in this work since the hydrogen bonds that keep bases together have a low binding energy.
The energetic analysis of these results show that 17.9 eV transferred during a central
impact would only produce base damage while keeping the DNA backbone intact. Current
Monte Carlo-based approaches to study the early DNA damage induced by ionizing radiation
15
a)b)c)d)xzyGUACYTGUAGUAGUACYTCYTCYTH+H+ leftcarrying charge1.21 fs2.42 fs100 fs193 fsFigure 5: System snapshots for some important stages during the proton-bound DNA upper
collision: a) the proton just passing through the bp, b) the proton leaves the bp taking
a fraction of the system charge (electron capture process), c) beginning of the dissociation
process, and d) the bp dissociates with into five fragments. Animation with the whole process
can be found in supplementary information.
16
xzya)b)c)d)CYTCYTCYTCYTGUAGUAGUAGUA1.21 fs58.1 fs100 fs193 fssuppose that an energy transfer to the phosphate-sugar group above a given threshold is
enough to induce a single strand break. This threshold is commonly set to about 10 eV.
Those approaches also consider that only the DNA component directly impacted by the
projectile is affected. Those components are commonly divided into the sugar-phosphate
groups and the nitrogeneous bases. This work shows that an impact on the phosphate group
with an energy transfer of about 60 eV would be enough to break the DNA back bone
on the impacted side and to damage the bases as well while keeping intact the opposite
sugar-phosphate group. Furthermore, it is observed, as expected, that the proton-DNA
collision is a many-body problem during which all atoms feel the proton impact and receive
a fraction of the transferred energy. Thus, there is no passive DNA base pair component
during the collision, unlike it is supposed on the vast majority of biophysical models based
on Monte Carlo simulations mentioned just above. A complete TD-DFT study of this
problem could provide enough information in order to change the actual paradigm of such
biophysical models. That is, DNA damage probabilities could be determined as a function
of the projectile impact parameter and several sites can be damage in a single impact, as
the backbone and bases.
This study also provides insights on the time evolution of the DNA base pair after the
proton impact. Dissociation times are consistent with those reported in the literature for
large molecules. The linking of the base pair to its neighbors tends to delay the dissociation
process, which is about 100 fs according to our results.
Conclusions
TD-DFT can be a useful tool to study early DNA damage induced by the impact of heavy
charged particle. However, due to the enormous computing resources and time demanded
even for a single DNA base pair, a systematic study of this process is difficult to accomplish.
A complete study should include the combination of many impact parameters, energies,
17
and incidence directions of the projectile. Yet, some interesting features can be obtained
even with a limited number of calculations. For instance, energy transfers required for DNA
damage can be inferred. In addition, some light can be shed on the effect of base-pair linking
to its neigbors and the importance of the impact site on the way DNA is damaged by the
ionizing particle. A relatively high energy transferred during the impact of a proton on the
phosphate-sugar group can induce a single strand break together with bases damage.
It
seems that every backbone break is accompanied by base damage. This is an important
point since current Monte Carlo-based approaches in computational radiobiology suppose
that base damage is independent of backbone break. A central impact with energy transfer
of less than 20 eV would not be enough to produce backbone breaks. Current biophysical
models used to study the early DNA damage induced by charged particles can be improved
with studies like this one.
Acknowledgements
M.B. thanks the Conselho Nacional para o Desenvolvimento Cient´ıfico e Tecnol´ogico (CNPq),
Brazil, for financing his research activities through the project 306775/2015-8. A.R. and
U.D.G. acknowledge financial support from the European Research Council(ERC-2015-AdG-
694097), Grupos Consolidados (IT578-13), H2020-NMP-2014 project MOSTOPHOS (GA
no. 646259), European Union's H2020 programme under GA no.676580 (NOMAD) and
COST Action MP1306 (EUSpec). The images of this work were made with VMD soft-
ware support. VMD is developed with NIH support by the Theoretical and Computational
Biophysics group at the Beckman Institute, University of Illinois at Urbana-Champaign.
References
(1) Sax, K. Chromosome aberrations induced by X-rays. Genetics 1938, 23, 494 -- 516.
18
(2) Lea, D. E.; Catcheside, D. G. The mechanism of the induction by radiation of chromo-
some aberrations inTradescantia. Journal of Genetics 1942, 44, 216 -- 245.
(3) Kellerer, A. M.; Rossi, H. H. The action of dual radiation action. Current Topics in
Radiation Rsearch Quarterly 1972, 8, 85 -- 158.
(4) Pomplun, E. A new DNA target model for track structure calculations and its first
application to I-125 Auger electrons. Int. J. Radiat. Biol. 1991, 59, 625 -- 642.
(5) Nikjoo, H.; O'Neil, P.; Terrissol, M.; Goodhead, D. T. Quantitative modelling of DNA
damage using Monte Carlo track structure method. Radiat. Environ. Biophys. 1999,
38, 31 -- 38.
(6) Friedland, W.; Jacob, P.; Kundr´at, P. Mechanistic simulation of radiation damage
to DNA and its repair: on the track towards systems radiation biology modelling.
Radiation Protection Dosimetry 2011, 143, 542 -- 548.
(7) Bernal, M. A.; deAlmeida, C. E.; David, M.; Pires, E. Estimation of the RBE of
mammography-quality beams using a combination of a Monte Carlo code with a B-
DNA geometrical model. Physics in Medicine and Biology 2011, 56, 7393 -- 7403.
(8) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Physical Review 1964, 136,
B864 -- B871.
(9) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation
Effects. Physical Review 1965, 140, A1133 -- A1138.
(10) Bacchus-Montabonel, M.-C.; Tergiman, Y. S. Radiation damage on biomolecular sys-
tems: Dynamics of ion induced collision processes. Computational and Theoretical
Chemistry 2012, 990, 177 -- 184.
(11) Werner, H. J.; Knowles, P. J. MOLPRO package of ab initio programs (version 1010.1).
19
(12) Sadr-Arani, L.; Mignon, P.; Abdoul-Carime, H.; Farizon, B.; Farizon, M.; Chermette, H.
DFT study of the fragmentation mechanism of uracil RNA base. Physical Chemistry
Chemical Physics 2012, 14, 9855 -- 9870.
(13) Sadr-Arani, L.; Mignon, P.; Abdoul-Carime, H.; Farizon, B.; Farizon, M.; Chermette, H.
Hydrogen release from charged fragments of the uracil cation followed by their frag-
mentation: A DFT study. Chemical Physics Letters 2013, 583, 165 -- 169.
(14) Sadr-Arani, L.; Mignon, P.; Chermette, H.; Douki, T. Theoretical and experimental
study of the fragmentation of protonated uracil. Chemical Physics Letters 2014, 605,
108 -- 114.
(15) Sadr-Arani, L.; Mignon, P.; Chermette, H.; Abdoul-Carime, H.; Farizon, B.; Farizon, M.
Fragmentation mechanisms of cytosine, adenine and guanine ionized bases. Physical
Chemistry Chemical Physics 2015, 17, 11813 -- 11826.
(16) L´opez-Tarifa, P.; Herv´edu Penhoat, M.-A.; Vuilleumier, R.; Gaigeot, M.-P.; Rothlis-
berger, U.; Tavernelli, I.; Le Padellec, A.; Champeaux, J.-P.; Alcam´ı, M.; Moretto-
Capelle, P. et al. Time-dependent density functional theory molecular dynamics sim-
ulation of doubly charged uracil in gas phase. Central European Journal of Physics
2014, 12, 97 -- 102.
(17) Abolfath, R. M.; van Duin, A. C. T.; Brabec, T. Reactive Molecular Dynamics Study
on the First Steps of DNA Damage by Free Hydroxyl Radicals. The Journal of Physical
Chemistry A 2011, 115, 11045 -- 11049.
(18) van Duin, A. C. T.; Dasgupta, S.; Lorant, F.; III, W. A. G. ReaxFF: A Reactive Force
Field for Hydrocarbons. The Journal of Physical Chemistry A 2001, 105, 9396 -- 9409.
(19) Abolfath, R. M.; Carlson, D. J.; Chen, Z. J.; Nath, R. A molecular dynamics simulation
of DNA damage induction by ionizing radiation. Physics in Medicine and Biology 2013,
58, 7143.
20
(20) Bernal, M. A. et al. Track structure modeling in liquid water: A review of the Geant4-
DNA very low energy extension of the Geant4 Monte Carlo simulation toolkit. Physica
Medica: European Journal of Medical Physics 2015, 31, 861 -- 874.
(21) Bottlander, D.; Mucksch, C.; Urbassek, H. M. Effect of swift-ion irradiation on DNA
molecules: A molecular dynamics study using the REAX force field. Nuclear Instru-
ments and Methods in Physics Research Section B: Beam Interactions with Materials
and Atoms 2015, 365, Part B, 622 -- 625.
(22) Surdutovich, E.; Yakubovich, A. V.; Solov'yov, A. V. Biodamage via shock waves ini-
tiated by irradiation with ions. Scientific Reports 2013, 3, 1289 EP -- .
(23) de Vera, P.; Mason, N. J.; Currell, F. J.; Solov'yov, A. V. Molecular dynamics study of
accelerated ion-induced shock waves in biological media. The European Physical Journal
D 2016, 70, 183.
(24) Bacchus-Montabonel, M.-C.; Calvo, F. Nanohydration of uracil: emergence of three-
dimensional structures and proton-induced charge transfer. Physical Chemistry Chem-
ical Physics 2015, 17, 9629 -- 9633.
(25) Bacchus-Montabonel, M. C.; Calvo, F. Influence of microhydration on the structures
and proton-induced charge transfer in RNA intermediates. Journal of Molecular Mod-
eling 2016, 22, 262.
(26) Miguel A.L. Marques, N. T. M., Nogueira, F. M., Gross, E., Rubio, A., Eds. Fundamen-
tals of Time-Dependent Density Functional Theory; Lecture Notes in Physics; Springer,
2012; Vol. 837.
(27) Protein Data Base. 2013; http://www.pdb.org/pdb/home/home.do.
(28) Castro, A.; Appel, H.; Oliveira, M.; Rozzi, C. A.; Andrade, X.; Lorenzen, F.; Mar-
ques, M. A. L.; Gross, E. K. U.; Rubio, A. octopus: a tool for the application of
21
time-dependent density functional theory. physica status solidi (b) 2006, 243, 2465 --
2488.
(29) Andrade, X.; Alberdi-Rodriguez, J.; Strubbe, D. A.; Oliveira, M. J. T.; Nogueira, F.;
Castro, A.; Muguerza, J.; Arruabarrena, A.; Louie, S. G.; Aspuru-Guzik, A. et al. Time-
dependent density-functional theory in massively parallel computer architectures: the
octopus project. Journal of Physics: Condensed Matter 2012, 24, 233202.
(30) Andrade, X. et al. Real-space grids and the Octopus code as tools for the development of
new simulation approaches for electronic systems. Physical Chemistry Chemical Physics
2015, 17, 31371 -- 31396.
(31) Perdew, J. P.; Zunger, A. Self-interaction correction to density-functional approxima-
tions for many-electron systems. Physical Review B 1981, 23, 5048 -- 5079.
(32) Troullier, N.; Martins, J. L. Efficient pseudopotentials for plane-wave calculations. Phys-
ical Review B 1991, 43, 1993 -- 2006.
(33) De Giovannini, U.; Larsen, A. H.; Rubio, A. Modeling electron dynamics coupled to
continuum states in finite volumes with absorbing boundaries. The European Physical
Journal B 2015, 88, 56.
(34) Humphrey, W.; Dalke, A.; Schulten, K. VMD: visual molecular dynamics. J Mol. Graph
1996, 14, 33 -- 8, 27 -- 8.
(35) Range, K.; McGrath, M. J.; Lopez, X.; York, D. M. The Structure and Stability of
Biological Metaphosphate, Phosphate, and Phosphorane Compounds in the Gas Phase
and in Solution. Journal of the American Chemical Society 2004, 126, 1654 -- 1665.
(36) Legon, A. C.; Millen, D. J. Angular geometries and other properties of hydrogen-bonded
dimers: a simple electrostatic interpretation of the success of the electron-pair model.
Chemical Society Reviews 1987, 16, 467 -- 498.
22
(37) Huheey, J.; Keiter, E. A.; Keiter, R. L. Inorganic Chemistry; Harper Collins, 1993.
(38) Gross, J. H. Mass spectrometry; Springer, 2011.
23
|
1210.1172 | 1 | 1210 | 2012-10-03T17:01:58 | Modeling self-organized systems interacting with few individuals: from microscopic to macroscopic dynamics | [
"physics.bio-ph",
"cs.SI",
"physics.soc-ph",
"q-bio.QM"
] | In nature self-organized systems as flock of birds, school of fishes or herd of sheeps have to deal with the presence of external agents such as predators or leaders which modify their internal dynamic. Such situations take into account a large number of individuals with their own social behavior which interact with a few number of other individuals acting as external point source forces. Starting from the microscopic description we derive the kinetic model through a mean-field limit and finally the macroscopic system through a suitable hydrodynamic limit. | physics.bio-ph | physics |
Modeling self-organized systems interacting with few individuals: from
microscopic to macroscopic dynamics
aUniversity of Ferrara, Department of Mathematics, Via Machiavelli 35, I-44121 Ferrara, Italy.
G. Albia, L. Pareschia
Abstract
In nature self-organized systems as flock of birds, school of fishes or herd of sheeps have to deal with
the presence of external agents such as predators or leaders which modify their internal dynamic.
Such situations take into account a large number of individuals with their own social behavior which
interact with a few number of other individuals acting as external point source forces. Starting
from the microscopic description we derive the kinetic model through a mean-field limit and finally
the macroscopic system through a suitable hydrodynamic limit.
Keywords: Kinetic models, mean field models, flocking, swarming, collective behavior
1. Introduction
The aim of this paper is to present different level of descriptions for the dynamic of a large group
of agents influenced by a small number of external agent. In a biological context this corresponds to
the behavior of a flock or a school of fishes attacked by one or more predators, or the movement of
a heard of sheep guided by a sheepdog. Recently such dynamics have been studied also in robotic
research, where engineers tried to control the action of a school of fishes introducing a fishbot
recognized as leader, see [4].
From the modeling viewpoint this turns out in considering a microscopic dynamic described by
classical flocking models like Cuker-Smale and D'Orsogna-Bertozzi et al. (see [10, 12]) interacting
with a set of few individuals characterized in an analogous way as in [9]. Moreover, motivated by
the works of [5], we endowed the classical dynamic of interaction both with a metric as well as a
topological interaction rule.
Following the approach in [8] we start from the microscopic dynamic, given by a ODEs system,
and we derive two other different level of description: the mesoscopic (or kinetic) level through a
mean-field limit and the macroscopic level through a suitable hydrodynamic limit. At difference
to the first order macroscopic models proposed in [9] here we obtain second order models for the
corresponding continuum dynamic.
Finally we report some numerical examples for the solution of the mean field model in a series
of test cases. The simulations have been performed using the fast algorithm recently presented
in [1]. Rigorous mathematical results concerning the asymptotic limits just described can be found
in [2].
Preprint submitted to Elsevier
August 10, 2018
2. Microscopic model
We are interested in the study of a dynamical system composed of N individuals and N p
external agents with the following general structure
xi = vi
i = 1, . . . , N
vi =
1
N ∗ Xj∈Σ∗
N (xi)
F (xi, vi, xj, vj) +
1
Np
Np
Xk=1
F p(xi, vi, pk)
(1)
ph = ϕh(t, p, AρN (ph))
h = 1, . . . , Np
where (x, v)i = (xi, vi) lives in R2d, d ≥ 1, i = 1, . . . , N and (p)h = ph ∈ Rnd, with n = 1, 2,
h = 1, . . . , Np, and Np ≪ N .
Function F describes the interactions inside the swarm, and F p depicts the interaction with
each external agent ph. According to the three zone model [3], F can be decomposed in
F (xi, vi, xj, vj) = H(xi, xj)(vj − vi) + A(xi, xj) + R(xi, xj) + S(vi)vi.
(2)
H characterize the alignment term, A the attraction, R the repulsion and S represent a self
propulsion-friction term. The same decomposition holds for F p if ph = (xp
h), i.e. n = 2.
In a first order model, n = 1, similar interaction can be considered, bu for the first term we should
speak of consensus dynamic rather then alignment.
h, vp
Moreover we endow the model with a topological rule of interaction. Each agent will interact
only with a fix number of agents of their specie,
Σ∗
N (xi) = {the N ∗ closest neighbors respect to i} .
A topological interaction is motivated by some recent studies, see [5, 11], suggesting that in a
flock each individual modifies its position according to few individuals directly surrounding it, no
matter how close or how far away those agents are. If N ∗ = N each agent interacts with all the
others and the topological interaction coincides with the global metric interaction [1, 7]. Functions
ϕh : [0, +∞) × (Rnd)Np × R −→ R, describe the evolution in time of each external individual and
they depend on the discrete density ρN defined as empirical measure
ρN (x, t) =
1
N
N
Xi=1
δ(x − xi(t)).
According to [9] we define A as a convolutional operator AρN (t) = ρN (t) ∗ η, where η is a smooth
kernel with compact support.
2.1. Classical swarming models
The classical swarming models take in account a global interaction between the agents, that
corresponds in our case to choose N ∗ = N neighbors.
2
Cucker-Smale model. Describes an alignment dynamic, see [10] and [7],
H(xi, xj)(vj − vi) = H(xi − xj )(vj − vi),
(3)
where H(xi − xj) is a function that measures the strength of the interaction between individuals
i and j , and depends on the mutual distance, under the assumption that closer individuals have
more influence than the far distance ones, as depicted by
H(r) =
1
(1 + r2)γ ,
where γ ≥ 0 discriminates the behavior of the solution, see [10, 7] and [8].
D'Orsogna, Bertozzi et al.. , considers a self-propelling, attraction and repulsion dynamic, see [12]
A(xi, xj ) + R(xi, xj) + S(vi)vi = −∇xiW (xj − xi) + (α − βvi2)vi
(4)
where W : Rd −→ R is a given potential modeling the short-range repulsion and long-range
attraction and α, β are positive parameters. A possible description is given by the following power
law
W (r) =
−
ra
a
rb
b
,
where a > b > 0 are positive parameters.
Both the models take in account symmetric interaction between agents, which correspond to
the conservation of the first momentum. This assumption isn't really realistic if we want to model
group of animals, to improve the models other features have been introduced as a perception
cone [1, 7] or the concept of relative distance [14]. In these cases we lose the interaction symmetry
and consequently the conservation of the first momentum, note that also the topological interaction
breaks the symmetry.
3. Kinetic model
In order to retrieve a mesoscopic description of the system (1) we proceed through a mean-field
limit, as already done in [2, 7, 8]. Since the limit is done only for the first set of equations, which
describe the evolution of the swarm, we obtain an hybrid model made by one kinetic equation and
the ODEs system governing the external agents.
The basic idea of the mean-field limit is to derive, through a weak argument, a single evolu-
tionary equation for f N , the empirical measures defined as
f N (x, v, t) =
1
N
N
Xi=1
δ(x − xi(t))δ(v − vi(t)).
A second step is to show that the limit f N −→ f for N −→ ∞ holds true, then rigorous derivation
of the limiting kinetic equation can be proved. A well-posedness theory for this derivation has been
developed in [7] for a general set of swarming models.
3
We remark that in the mean-field limit the analogous of the topological set, Σ∗
N (x), of interaction
is described by a characteristic function on the ball B(x, R∗), with center x and radius R∗ such
that
The model at mesoscopic level reads
R > 0, ρ∗ =ZRdZB(x,R)
f (x, v)dxdv) .
R∗ = min(R(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
∂tf + v · ∇xf = −∇v · (E ∗(x, v)f ) − ∇v · (E p(x, v)f ).
ph = ϕh(t, p, Aρ(ph))
h = 1, . . . , Np
(5)
where ρ(x, t) =RRd f (x, v, t)dv.
Remark 1. If we consider the decompositions (2) in the case of Cucker-Smale and D'Orsogna-
Bertozzi et al. models and ph = (xp
h), we have a more particular formulation for E ∗ and E p, in
this way
h, vp
E ∗(x, v) =
and
1
ρ∗ ZRdZB(x,R∗)
ρ∗ ZB(x,R∗)
−
1
H(x − y)(w − v)f (y, w)dydw+
∇xW (x − y)ρ(y)dy + S(v)v
E p(x, v) =
1
Np
Np
Xk=1
4. Hydrodynamic model
H p(x − xp
k)(vp
k − v) −
1
Np
Np
Xk=1
∇xW p(x − xp
k)
(6)
(7)
We are interested also in the macroscopic description of the system, from a numerical point of
view this corresponds to reduce the dimensionality of the problem in such a way that simulations
become affordable. As done in [2, 8, 13] we define
ρ := ρ(x, t) =Z f dv,
ρu := ρu(x, t) =Z vf dv, T := T (x, t) =Z v − u2f dv,
in order to obtain a system of equations which describe the evolution of these quantities, we
integrate the kinetic equation in (5) against dv and vdv.
According to [8] we impose the closure the momentum assuming that the fluctuations are
negligible, i.e., that the temperature T (x, t) = 0, and the velocity distribution is monokinetic:
f (x, v, t) = ρ(x, t)δ(v − u(x, t)). The previous assumptions lead to the following hydrodynamic
system
∂tρ + divx(ρu) = 0,
∂t(ρu) + ∇x · (ρu ⊗ u) = F ∗(x, u)ρ(x, t) + F p(x, u)ρ(x, t),
(8)
ph = ϕh(t, p, Aρ(ph))
h = 1, . . . , Np
4
4
3
2
1
Y
0
−1
−2
−3
−4
−4
t=0.1
−3
−2
−1
0
X
1
2
3
4
0.5
0.45
0.4
0.35
0.3
4
3
2
1
0.25
Y
0
0.2
0.15
0.1
0.05
0
−1
−2
−3
−4
−4
t=0.8
−3
−2
−1
0
X
1
2
3
4
4
3
2
1
Y
0
−1
−2
−3
−4
−4
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
t=4.4
−3
−2
−1
0
X
1
2
3
4
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
Figure 1: Shepherd dogs. With parameters a = 2.5, b = 0.1, γ = 0.45 for models (3) and (4).
In the particular case of Cucker-Smale and D'Orsogna Bertozzi et al. model and ph = (xp
second term reads
h, vp
h) the
F ∗(x, u) =
1
1
−
ρ∗ ZB(x,R∗)
ρ∗ ZB(x,R∗)
Xk=1
Np
F p(x, u) =
1
Np
5. Numerical examples
H(x − y)(u(y, t) − u(x, t))ρ(y, t)dy+
∇xW (x − y)ρ(y, t)dy + S(u)u
(9)
H p(x − xp
k)(vp
k − u(x, t)) −
1
Np
Np
Xk=1
∇xW p(x − xp
k)
(10)
We show here two numerical examples taken from [9] for two different biological dynamics. In
order to solve the kinetic part of the model, the simulations are performed using the Asymptotic
Binary Interaction algorithms proposed in [1]. We recall that the algorithm is based on a stochastic
routine which use Ns sample particles. The overall cost is O(Ns).
The dynamics considered for the swarm in both the cases are
F (xi, vi, xj, vj) = H(xi − xj)(vj − vi) + ∇xiW (xi − xj),
and only a repulsion dynamic respect to ph in F p, given by the W p(r) = −rc/c with c > 0, and
where r = xi − ph.
Shepherd dogs. The simulation represents the evolution of a swarm controlled by two leaders p1, p2 ∈
R2n, n = 1, interacting with the swarm in the following way
ϕh(t, p, sh) = Vp
; Vp = 300,
rp = 5,
h = 1, 2.
η(x) =
3
πr6
p
(max{0, r2
p − x2})2,
sh := Aρ(xp
h) = (ρ ∗x ∇η)(xp
h).
s⊥
h
p1 + sh2
5
Y
5
4
3
2
1
0
−1
−2
−5
t=0.3
−4
−3
−2
−1
0
X
1
2
3
4
5
Y
7
6
5
4
3
2
1
0
−1
−2
0.2
0.15
0.1
0.05
0
t=0.4
t=0.5
Y
7
6
5
4
3
2
1
0
−1
−2
0.12
0.1
0.08
0.06
0.04
0.02
0
−6
−4
−2
0
X
2
4
6
−6
−4
−2
0
X
2
4
6
0.11
0.1
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01
Figure 2: Swarm attacked by a predator. With parameters a = 4, b = 2, γ = 0.45 for models (3) and (4).
Swarm attacked by predator. We consider the evolution of a swarm which undergoes the action of
a predator. The predator is modeled by the evolution of p = (xp, vp) ∈ R2n, n = 2 and its evolution
is lead by the following potential
ϕ(t, p, s) = (vp, Vps); Vp = 1500,
rp = 5.
η(x) =
3
πr6
p
(max{0, r2
p − x2})2,
s := Aρ = ρ ∗ ∇η.
References
[1] G. Albi and L. Pareschi. Binary interaction algorithms for the simulations of flocking and swarming
dynamics. SIAM, Multiscale Modeling and Simualtion, to appear, 2012.
[2] G. Albi and L. Pareschi. Kinetic and hydrodynamic models for self-organized systems interacting with
few individuals. preprint, 2012.
[3] I. Aoki. A simulation study on the schooling mechanism in fish. Bulletin Of The Japanese Society Of
Scientific Fisheries, 48(8):1081 -- 1088, 1982.
[4] M. Aureli, F. Fiorilli, M. Porfiri. Portraits of self-organization in fish schools interacting with robots.
Physica D, Vol. 241, Issue 9, pages 908 -- 920, May 2012.
[5] M. Ballerini, N. Cabibbo, R. Candelier, A. Cavagna, E. Cisbani, I. Giardina, V. Lecomte, A. Orlandi,
G. Parisi, A. Procaccini, et al. Interaction ruling animal collective behavior depends on topological
rather than metric distance: Evidence from a field study. Proceedings of the National Academy of
Sciences, 105(4):1232, 2008.
[6] J. A. Canizo , J. A. Carrillo, J. Rosado. A well-posedness theory in measures for some kinetic models of
collective motion. Mathematical Models and Methods in Applied Sciences, Vol. 21, No. 3, pp. 515-539,
March 2011.
[7] J. A. Carrillo, M. Fornasier, J. Rosado, and G. Toscani. Asymptotic flocking dynamics for the kinetic
Cucker-Smale model. SIAM J. Math. Anal., 42(1):218 -- 236, 2010.
6
[8] J.A. Carrillo, M. Fornasier, G. Toscani, and F. Vecil. Particle, kinetic, and hydrodynamic models of
swarming. pages 297 -- 336, 2010.
[9] R. Colombo and M. L´ecureux-Mercier, An Analytical Framework to Describe the Interactions Between
Individuals and a Continuum. Journal of Nonlinear Science, 22, pages 39 -- 61, 2012.
[10] F. Cucker and S. Smale. Emergent behavior in flocks. IEEE Trans. Automat. Control, 52(5):852 -- 862,
2007.
[11] E. Cristiani, B. Piccoli, and A. Tosin. Modeling self-organization in pedestrians and animal groups
from macroscopic and microscopic viewpoints.
In Mathematical modeling of collective behavior in
socio-economic and life sciences, Model. Simul. Sci. Eng. Technol., pages 337 -- 364. Birkhauser Boston
Inc., Boston, MA, 2010.
[12] M.R. D'Orsogna, Y.L. Chuang, A.L. Bertozzi, and L.S. Chayes. Self-propelled particles with soft-core
interactions: patterns, stability, and collapse. Physical review letters, 96(10):104 -- 302, 2006.
[13] S. Ha and E. Tadmor. From particle to kinetic and hydrodynamic descriptions of flocking. Kinet. Relat.
Models, 1(3):415 -- 435, 2008.
[14] S. Motsch and E. Tadmor. A new model for self-organized dynamics and its flocking behavior. Journal
of Statistical Physics, Springer, 144(5):923-947, 2011.
7
|
1306.4942 | 1 | 1306 | 2013-06-18T10:43:16 | Disentangling electronic and vibronic coherences in two-dimensional echo spectra | [
"physics.bio-ph",
"physics.chem-ph"
] | The prevalence of long-lasting oscillatory signals in the 2d echo-spectroscopy of light-harvesting complexes has led to a search for possible mechanisms. We investigate how two causes of oscillatory signals are intertwined: (i) electronic coherences supporting delocalized wave-like motion, and (ii) narrow bands in the vibronic spectral density. To disentangle the vibronic and electronic contributions we introduce a time-windowed Fourier transform of the signal amplitude. We find that 2d spectra can be dominated by excitations of pathways which are absent in excitonic energy transport. This leads to an underestimation of the life-time of electronic coherences by 2d spectra. | physics.bio-ph | physics |
Disentangling electronic and vibronic coherences in two-dimensional echo spectra
Christoph Kreisbeck,1 Tobias Kramer,2, 3 and Al´an Aspuru-Guzik1
1Department of Chemistry and Chemical Biology,
Harvard University, Cambridge, Massachusetts 02138, USA
2Institut fur Physik, Humboldt Universitat zu Berlin, 12489 Berlin, Germany
3Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA
(Dated: August 13, 2018)
The prevalence of long-lasting oscillatory signals in the 2d echo-spectroscopy of light-harvesting
complexes has led to a search for possible mechanisms. We investigate how two causes of oscillatory
signals are intertwined: (i) electronic coherences supporting delocalized wave-like motion, and (ii)
narrow bands in the vibronic spectral density. To disentangle the vibronic and electronic contri-
butions we introduce a time-windowed Fourier transform of the signal amplitude. We find that 2d
spectra can be dominated by excitations of pathways which are absent in excitonic energy transport.
This leads to an underestimation of the life-time of electronic coherences by 2d spectra.
I.
INTRODUCTION
The observation of long-lasting oscillations in peak
amplitudes of two-dimensional (2d) echo spectroscopy
as function of increasing delay time between pump and
probe pulses1 -- 6 has lead to an intensive search for
quantum mechanical effects in photosynthetic systems
at physiological temperatures. The general theoretical
model of excitonic energy transfer from the antenna to
the reaction center is well established.7 -- 9 For the Fenna-
Matthews-Olson (FMO) complex an efficient transfer re-
lies on a dissipative coupling of electronic and vibronic
degrees of freedom,10 -- 16 which puts the excited system
in an energy funnel towards the reaction center. The
oscillations recorded in 2d echo spectra were not antici-
pated within a simplified rate equation formalism based
on Markovian approximations for the dissipative coupling
of vibronic and electronic degrees of freedom.17,18
The physical origin of the experimental observations
has lead to various theoretical proposals, which are
roughly divided into two mechanisms. As shown from
calculations in model Hamiltonians19 -- 21 the presence of
δ-spiked peaks in the vibronic mode distribution does
imprint a long-lasting oscillatory signal on coherences.
The spectral density determined from fluorescence line-
narrowing experiments (JWendling in Fig. 1) reveals a se-
ries of peaks. However a calculation of 2d spectra and
coherences22 using a close approximation to the density
(denoted by J3peaks and J11peaks in Fig. 1) puts the vi-
bronic contributions on a much smaller scale compared
to electronic coherences, which lead to oscillation periods
determined by the differences in excitonic eigenenergies.
The second proposed mechanism22 identifies two prereq-
uisites for long-lasting electronic coherences and oscilla-
tions of cross-peaks of 2d-spectra in the continuous part
of the vibronic mode distribution: (i) a small initial slope
of the vibronic density towards zero frequency, and (ii) a
larger coupling between a continuum of vibrations whose
frequencies are in the range of the typical excitonic en-
ergy gaps. Both conditions are fullfilled by the florescence
line-narrowing spectral density JWendling.
☛❲❡☞✌✍✎☞✏
☛✠✠✑❡✒✓✔
☛✕✑❡✒✓✔
✷✺✵
✷✵✵
✶✺✵
✶✵✵
✺✵
✵
✶✵✵
✷✵✵
✸✵✵
✖❤☎ ✆✝✞✟✠✡
✹✵✵
✺✵✵
FIG. 1. Spectral density of the FMO complex. Circles: mea-
sured spectral density,23 parametrization.24 Fits J{3,11}peaks
are used for the GPU-HEOM calculation. The marks on the
frequency axis indicate differences of exciton eigenenergies.
It remains an open question to what extend 2d spectra
are probing vibrational or electronic properties of light-
harvesting complexes. Moreover, for the design of artifi-
cial nanostructure it is of interest to identify the physi-
cal mechanism for efficient transport and conditions for
the accompanying vibronic density. Here, we disentan-
gle the interplay of electronic and vibronic effects on the
transport by introducing a reliable method to dissect the
2d-echo signal. We show that a time-windowed short-
time Fourier transform (STFT) is a suitable tool to track
and assign specific vibronic and electronic contributions
at all delay times. We find that two-dimensional echo-
spectra can significantly underestimate the duration of
electronic coherences due to the destructive superposi-
tion of several pathways related to additional excitations
and ground-state bleaching. Such pathways are not im-
portant for excitonic energy transfer, which for the same
system displays longer electronic coherences than the 2d
spectra. The connection between 2d spectra and efficient
❏
✭
✦
✮
✭
✁
✂
✄
✮
✁✄
✁✵
✁✂
✲ ✁✂
t✓
t✔
❘✌✍✚❊✶❊✺ ❪
✎✏✑✒ ❋✇
✖❤✦ ❬✄☎✆✶❪
✖❤✦✶
2
✖❤✡❦
✁☎
✂
✂✁☎
t ❬✡☛☞
FIG. 3. STFT analysis (tw = 0.4 ps, β = 7) of the coher-
ence ρE1E5 of Fig. 2. The left arrow marks the frequency of
the electronic energy difference E1 − E5 and the right arrows
indicate the centers of vibrational peaks of J11peaks, Fig. 1.
✝✞ ❬✟✠❪
FIG. 2. Time evolution of the coherence ρE1E5 (t) for the
spectral density J11peaks at T = 150 K. The dashed line shows
a typical window function (tc = 0.6 ps, tw = 0.4 ps, β = 7).
transport and the role of coherences requires a careful
evaluation.
II. TRANSITION FROM ELECTRONIC TO
VIBRATIONAL INDUCED COHERENCE IN THE
FMO COMPLEX
Multiple frequencies are a common pattern in the
cross-peak oscillations observed in 2d echo spectra of var-
ious LHCs such as the FMO complex25,26 or light harvest-
ing proteins PE545 and PC645 from marine cryptophyte
algae.3,6 To unravel which physical processes are present
as the delay-time progresses and to uncover the tran-
sition from electronic to vibrational induced coherences
we introduce a windowed short-time Fourier transform
(STFT) of the signal
Fw(ω, tc) ρEiEk (t) = Z ∞
−∞
dt Fw(tc, t)ρEiEk (t)eiωt
(1)
where tc denotes the center and tw the width of the win-
dow function
Fw(tc, t) = Xs=±1
s
1 + eβ(t−tc−stw/2)/tw
.
(2)
Later, we normalize Fw(tc)ρEiEk to its maximal value.
A typical window function along with the coherence be-
tween exciton eigenstates E1i and E5i of the FMO com-
plex calculated in Ref.22 is shown in Fig. 2 together with
the corresponding STFT analysis in Fig. 3. Sweeping
the centre of the window function tc yields information
about the changes in the frequency distribution with ad-
vancing time in the exciton dynamics.
In Fig. 3, the
initially dominant electronic coherent frequency ¯hω15 is
indicated by the arrow on the left side. At later times
dephasing reduces the influence of the electronic coher-
ence while the interaction with the strongly coupled vi-
brational modes increases. Around tc = 0.7 ps a "gap"
can be seen, where the electronic frequency is absent but
shortly afterwards re-emerges. From 1 − 3.5 ps a wob-
bling motion of the dominant contribution is seen, which
is caused by the vibronic frequencies marked by arrows
on the right hand side. During the time evolution, en-
ergy is shared among different vibrational modes close in
resonance with the exction frequency ¯hω15. Which vibra-
tions are activated changes with time. For example, the
vibration at ¯hΩ = 380 cm−1 is present around 2.2 ps but
absent around 3 ps. The features revealed in the STFT
reflect two different mechanisms in the energy transfer.
Electronic coherence signifies delocalized wavelike energy
transfer which helps to overcome energy barriers and fa-
cilitates fast energy transfer to specific target states. On
the other hand the signatures of strongly coupled vibra-
tional modes mark reversible energy exchange between
the exciton system and the protein environment.27
III. COHERENCES IN 2D SPECTRA
The decay of peak amplitude oscillations in two-
dimensional spectra has been used to extract coherence
lifetimes.5 In the following, we identify two processes
which reduce the signature of electronic coherence in
two-dimensional spectra compared to transport calcula-
tions. To elucidate and quantify the interplay between
electronic and vibronic degrees of freedom we discuss
the physical processes mapped by 2d-echo spectra within
a model trimer consisting of sites 1 − 3 of the FMO
complex24
Hex =
410 −87.7 5.5
30.8
210
530
30.8
−87.7
5.5
cm−1.
(3)
❝
✆
✝
✞
✟
✞
✠
❝
✞
◆
✭
t
❝
✮
✁
✂
❋
✇
✸✵✵
✷✵✵
✶✵✵
✵
✵
✶✵✵
✷✵✵
✸✵✵
✖❤☎ ✆✝✞✟✠✡
✹✵✵
✺✵✵
FIG. 4. Five peak model spectral density J5peaks with a
suppressed slope at zero frequency and three strongly cou-
pled underdamped vibrational modes. The vibrational fre-
quencies (dashed line) are close in resonance with the dif-
ference of the exciton energies Ei − Ej (solid line) of the
exciton system Eq. (3). Parameters for Eq. (4): λk ∈
{11, 9, 12, 7, 4} cm−1, ν −1
k ∈ {230, 50, 750, 700, 750} fs, ¯hΩk ∈
{80, 260, 135, 260, 430} cm−1.
As for the FMO complex22 we describe the energy trans-
fer within a Frenkel exciton model,28 linearly coupled to
independent baths at each site. The spectral density is
parametrized by a superposition of shifted Drude-Lorentz
peaks
J(ω) =
M
Xk=1
(cid:20)
νkλkω
k + (ω + Ωk)2 +
ν2
νkλkω
k + (ω − Ωk)2(cid:21) .
ν2
(4)
The spectral density J5peaks (Fig. 4) for the model trimer
fulfils the criteria for long-lived electronic coherences due
to its small initial slope and later on larger coupling.
To study the role of specific vibrational modes J5peaks
adds three vibrational modes centered at frequencies
¯hΩ1 = 135 cm−1, ¯hΩ2 = 260 cm−1, ¯hΩ3 = 430 cm−1,
which puts them close to resonance with the differences
in exciton eigenenergies. The total 2d-echo rephasing
signal is composed of three Liouville pathways29,30 re-
flecting stimulated emission (SE), ground-state bleach-
ing (GB) and excited state absorption (ESA), where the
laser pulses create two excitons in the system. Formally,
we calculate the Fourier transform of the third order re-
sponse function SRP (t3, Tdelay, t1)29,30 in the impulsive
limit
dt1dt3eiω3t3−iω1t1
0
RP (t3, Td, t1) + SESA
(5)
IRP (ω1, Td, ω3) = Z ∞
0 Z ∞
RP (t3, Td, t1) + SSE
×[SGB
RP (t3, Td, t1)]
We proceed along the steps in Ref.31 and include the ro-
tational average over random orientations by sampling
20 distinct laser polarization vectors aligned along the
vertices of a dodecahedron. We assume equal dipole
strength of the three pigments and dipole orientations
3
along the nitrogen atoms NB − ND.32 The propagation
is performed with the GPU-HEOM method.16,33 In an
experimental setup all three pathways are tied together
and cannot be analysed separately, while the theoreti-
cal calculations allows us to study the individual contri-
butions. Typically, the SE and ESA pathways interfere
destructively and lead to a diminished amplitude of the
coherent signal in the total spectrum. The cancellation of
coherent amplitudes in 2d spectra enhances the influence
of strongly coupled vibrations on the cross-peak oscilla-
tions.
In particular the ground-state vibrations moni-
tored by the GB get amplified in weight in the 2d-echo
spectra. Our finding, that ground-state vibrations af-
fect the cross-peak dynamics is in agreement with Tiwari
et al.34 who investigate the effect of a single vibrational
mode in resonance with a model dimer. Different condi-
tions govern transport and 2d-echo spectra, where sev-
eral pulses hit the sample. For instance the ground-state
bleaching and excited state absorption induced by the
2d-echo setup are not relevant for the energy transfer
process. Thus the apparent diminishing of coherences in
2d spectra does not imply unimportance of coherence in
the excitonic transfer.
The cancellation effects are prominently visible in Fig. 5,
which displays the oscillatory component of the lower-
diagonal cross-peak dynamics of CP(12), CP(13) and
CP(23) of the rephasing signal as function of delay time
Tdelay. The amplitude of CP(ik) is determined by in-
tegrating the real part of the rephasing 2d-spectra (all
three rephasing pathways) over a small rectangle (∆E =
36 cm−1) in the 2d-energy grid. The center of the rect-
angular area is located at ωi = Ei + λ/4.5 and ωk =
Ek + λ/4.5, where Ei, Ek denote the exciton eigenener-
gies of Eq. (3). The shift λ/4.5 takes into account the
stokes shift and ensures that the amplitude is averaged
around the maximum of the peaks. The non-oscillatory
background is fitted by f (t) = a + b e−ct + d t + d2 t2 and
is subtracted from the total signal. The three cross peaks
show different behaviours. In the upper panel CP(12) os-
cillates dominantly with a single frequency that matches
the vibrational mode ¯hΩ1. The 2d-spectra thus appar-
ently shows that there is almost no contribution of elec-
tronic coherence between exciton states E1i and E2i
left. This finding is in contradiction to simulations of
the coherence ρE1E2(t) that predict long lasting elec-
tronic coherence up to 1 ps. Studying the dynamics
of the SE and ESA pathways of CP(12) separately re-
covers the long lasting electronic coherence. But in the
case of CP(12) the amplitudes of the coherent oscilla-
tions of the SE and ESA pathway are nearly identical
and the destructive interference of these pathways can-
cels the contribution of electronic coherence in the total
signal. The remaining oscillatory component of the to-
tal signal reflects the ground state vibrations of the GB
pathway shown as dashed line in Fig. 5(a) and the signal
of electronic coherence is masked by the vibrational mode
¯hΩ1. The cross-peaks CP(13) and CP(23) (Fig. 5(b) and
Fig. 5(c)) show a more complex dynamics, which is best
❏
✭
✦
✮
❬
✁
✂
✄
❪
4
✖❤✡
✖❤✡✷
✖❤✡✶
✏✓✒
✖❤✡
✖❤✡✷
✖❤✡✶
✏✔✒
✖❤✡
✖❤✡✷
✖❤✡✶
(a)
(b)
(c)
1
0
-1
1
0
-1
1
0
]
.
u
.
a
[
P
C
]
.
u
.
a
[
P
C
]
.
u
.
a
[
P
C
Re{CP(12)}
only GB Re{CP(12)}
✖❤✦ ❬✝✞✟✶❪
❙☞✌☞❙✍
✏✑✒
SE+GB+ESA
Re{CP(13)}
SE+GB+ESA
Re{CP(23)}
✖❤✦✷
✖❤✦ ❬✝✞✟✶❪
●✎
✖❤✦✷
t❝ ❬✠☛❪
SE+GB+ESA
t❝ ❬✠☛❪
✖❤✦ ❬✝✞✟✶❪
-1
0
0.5
1
1.5
2
❙☞✌●✎✌☞❙✍
Tdelay [ps]
✖❤✦✷
FIG. 5. Oscillatory component of the real part of the cross-
peak dynamics of the total rephasing 2d-echo signal for Eq. (3)
and J5peaks as function of delay time Tdelay. The temperature
is set to T = 150 K.
analysed using the STFT methodology for the case of
CP(23) in Fig. 6. Multiple frequencies originating from
electronic coherence and vibronic modes superpose each
other and form rich structures in the beating pattern.
The STFT frequency analysis for the sum of ESA and SE
pathways is depicted in Fig. 6(a). In agreement with the
coherences (not shown) we observe a transition around
Tdelay = 0.7 ps from initially electronic coherence with
frequency ¯hω23 = 212 cm−1 to vibrational frequencies
¯hΩk at later times. The strongest coupling to the exciton
dynamics stems from the vibrational modes ¯hΩ1 and ¯hΩ2
that are close in resonance with the electronic frequency
¯hω23. For the GB pathway (Fig. 6(b)) the exciton system
remains in the electronic ground state during Tdelay and
no electronic frequencies show up in the cross peak dy-
namics. In the total signal Fig. 6(c) the GB vibrational
modes leave their trace and diminish the relative weight
of the electronic coherences.
t❝ ❬✠☛❪
FIG. 6. STFT analysis (tw = 0.7 ps, β = 7) of the cross-peak
dynamics CP(23) of the rephasing 2d-echo signal. Shown are
results for (a) the sum of the SE and ESA pathway, (b) the
contribution of the GB pathway and (c) the total signal in-
cluding all three pathways. The width of the window func-
tion is tw = 1.0 ps and T = 150 K. The left arrow marks
the frequency of electronic coherence ¯hω23 and the right ar-
rows mark the frequencies of the strongly coupled vibrational
modes ¯hΩk.
IV. CONCLUSION
The STFT provides a tool to determine the physical
mechanisms behind oscillatory signals for progressing de-
lay time. We find that the determination of electronic co-
herence life-times from the experimentally observed total
signal in the 2d-echo spectra is difficult and can underes-
timate the lifetime of electronic coherences. For varying
delay times the STFT shows multiple frequencies and a
mix of vibronic and electronic contributions. Similar ob-
◆
✭
❚
❞
✁
✂
✄
②
✮
☎
✆
❋
✇
◆
✭
❚
❞
✁
✂
✄
②
✮
☎
✆
❋
✇
◆
✭
❚
❞
✁
✂
✄
②
✮
☎
✆
❋
✇
servations are reported for experimentally recorded 2d
spectra.3,6,25,26 We further found examples of a reemer-
gent amplitude in the oscillatory component at later time
in CP(23), where a larger amplitude at Tdelay = 0.35 ps
is seen compared to Tdelay = 0.2 ps. Interestingly, such
reemerging amplitudes are also observed for the FMO
complex.25,26
The pollution of the 2d-signal by ground-state vibra-
tions and two-exciton states is not directly relevant for as-
sessing the role of coherences in excitonic energy transfer,
where these two pathways are absent. This has also im-
plications for identifying vibronic spectral-densities sup-
porting efficient transport. We find a good correspon-
dence between coherence lifetimes of the stimulated emis-
sion pathway (SE) of cross peak CP(i, j) and the corre-
sponding coherence ρEiEj . For certain cross-peaks the
5
ground state vibrations dominate, and the 2d-echo spec-
tra are not able to determine the electronic coherence
lifetime.
like the
proposed witness of electronic coherence35 or quantum
process tomography36,37 is required to reveal the hidden
information.
In this case an additional analysis,
ACKNOWLEDGMENTS
A.A.-G. and C.K. are supported by the DARPA grants
N66001-10-1-4063, N66001-10-1-4059, and T.K. by a
Heisenberg fellowship of the DFG (KR 2889/5). A.A.-G.
thanks the Corning foundation for their generous sup-
port.
1 Engel, G. S.; Calhoun, T. R.; Read, E. L.; Ahn, T.-
K.; Mancal, T.; Cheng, Y.-C.; Blankenship, R. E.; Flem-
ing, G. R. Evidence for wavelike energy transfer through
quantum coherence in photosynthetic systems. Nature
2007, 446, 782.
2 Lee, H.; Cheng, Y.-C.; Fleming, G. R. Coherence Dynam-
ics in Photosynthesis: Protein Protection of Excitonic Co-
herence. Science 2007, 316, 1462.
3 Collini, E.; Wong, C. Y.; Wilk, K. E.; Curmi, P. M. G.;
Brumer, P.; Scholes, G. D. Coherently wired light-
harvesting in photosynthetic marine algae at ambient tem-
perature. Nature 2010, 463, 644.
4 Calhoun, T. R.; Ginsberg, N. S.; Schlau-Cohen, G. S.;
Cheng, Y.-C.; Ballottari, M.; Bassi, R.; Fleming, G. R.
Quantum Coherence Enabled Determination of the En-
ergy Landscape in Light-Harvesting Complex II. J. Phys.
Chem. B 2009, 113, 16291.
5 Panitchayangkoon, G.; Hayes, D.; Fransted, K. A.;
Caram, J. R.; Harel, E.; Wen, J.; Blankenship, R. E.; En-
gel, G. S. Long-lived quantum coherence in photosynthetic
complexes at physiological temperature. Proc. Natl. Acad.
Sci. 2010, 107, 12766.
6 Turner, D. B.; Dinshaw, R.; Lee, K.-K.; Belsley, M. S.;
Wilk, K. E.; Curmi, P. M. G.; Scholes, G. D. Quan-
titative investigations of quantum coherence for a light-
harvesting protein at conditions simulating photosynthe-
sis. Phys. Chem. Chem. Phys. 2012, 14, 4857.
7 Amerongen, H. V.; Valkunas, L.; van Grondelle, R. Pho-
tosynthetic excitons; World Scientific: Singapore, 2000.
8 Blankenship, R. E. Molecular mechanisms of photosynthe-
sis; Blackwell Science: London, 2002.
9 Frigaard, N.-U.; Bryant, D. A. Seeing green bacteria in a
new light: Genomics-enabled studies of the photosynthetic
apparatus in green sulfur bacteria and filamentous anoxy-
genic phototrophic bacteria. Arch. Microbiol. 2004, 182,
265.
10 Mohseni, M.; Rebentrost, P.; Lloyd, S.; Aspuru-Guzik, A.
Environment-assisted quantum walks in photosynthetic
energy transfer. J. Chem. Phys. 2008, 129, 174106.
11 Rebentrost, P.; Mohseni, M.; Kassal, I.; Lloyd, S.; Aspuru-
Guzik, A. Environment-assisted quantum transport. New
J. Phys. 2009, 11, 033003.
12 Plenio, M. B.; Huelga, S. F. Dephasing-assisted transport:
Quantum networks and biomolecules. New. J. Phys. 2008,
10, 113019.
13 Caruso, F.; Chin, A. W.; Datta, A.; Huelga, S. F.; Ple-
nio, M. B. Highly efficient energy excitation transfer in
light-harvesting complexes: The fundamental role of noise-
assisted transport. J. Chem. Phys. 2009, 131, 105106.
14 Wu, J.; Liu, F.; Shen, Y.; Cao, J.; Silbey, R. J. Efficient en-
ergy transfer in light-harvesting systems, I: Optimal tem-
peratrue, reorganization energy, and spatial-temporal cor-
relations. New J. Phys. 2010, 12, 105012.
15 Mohseni, M.; Shabani, A.; Lloyd, S.; Rabitz, H. Optimal
and robust energy transport in light-harvesting complexes:
(II) A quantum interplay of multichromophoric geometries
and environmental interactions. arXiv:1104.4812v1 2011,
16 Kreisbeck, C.; Kramer, T.; Rodr´ıguez, M.; Hein, B. High-
Performance Solution of Hierarchical Equations of Motion
for Studying Energy Transfer in Light-Harvesting Com-
plexes. J. Chem. Theory Comput. 2011, 7, 2166.
17 Cho, M.; Vaswani, H. M.; Brixner, T.; Stenger, J.; Flem-
ing, G. R. Exciton Analysis in 2D Electronic Spectroscopy.
J. Phys. Chem. B 2005, 109, 10542.
18 Brixner, T.; Stenger, J.; Vaswani, H. M.; Cho, M.;
Blankenship, R. E.; Fleming, G. R. Two-dimensional spec-
troscopy of electronic couplings in photosynthesis. Nature
2005, 434, 625.
19 Christensson, N.; Kauffmann, H. F.; Pullerits, T.; Man-
cal, T. Origin of Long Lived Coherences
in Light-
Harvesting Complexes. J. Phys. Chem. B 2012, 116, 7449.
20 Chin, A.; Prior, J.; Rosenbach, R.; Caycedo-Soler, F.;
Huelga, S.; Plenio, M. The role of non-equilibrium vibra-
tional structures in electronic coherence and recoherence in
pigment-protein complexes. Nature Physics 2013, 9, 113 --
118.
21 Kolli, A.; O'Reilly, E. J.; Scholes, G. D.; Olaya-Castro, A.
The fundamental role of quantized vibrations in coher-
ent light harvesting by cryptophyte algae. The Journal of
Chemical Physics 2012, 137, 174109.
22 Kreisbeck, C.; Kramer, T. Long-Lived Electronic Coher-
ence in Dissipative Exciton Dynamics of Light-Harvesting
Complexes. J. Phys. Chem. Lett. 2012, 3, 2828.
23 Wendling, M.; Pullerits, T.; Przyjalgowski, M. A.;
Vulto, S. I. E.; Aartsma, T. J.; van Grondelle, R.;
van Amerongen, H. Electron-Vibrational Coupling in the
Fenna-Matthews-Olson Complex of Prosthecochloris aes-
tuarii Determined by Temperature-Dependent Absorption
and Fluorescence Line-Narrowing Measurements. J. Phys.
Chem. B 2000, 104, 5825.
24 Adolphs, J.; Renger, T. How Proteins Trigger Excitation
Energy Transfer in the FMO Complex of Green Sulfur Bac-
teria. Biophys. J. 2006, 91, 2778.
25 Hayes, D.; Wen, J.; Panitchayangkoon, G.; Blanken-
ship, R. E.; Engel, G. S. Robustness of electronic coher-
ence in the Fenna-Matthews-Olson complex to vibronic
and structural modifications. Faraday Discuss. 2011, 150,
459.
26 Panitchayangkoon, G.; Voronine, D. V.; Abramavicius, D.;
Caram, J. R.; Lewis, N. H. C.; Mukamel, S.; Engel, G. S.
Direct evidence of quantum transport in photosynthetic
light-harvesting complexes. Proc. Natl. Acad. Sci. 2011,
108, 20908.
27 Rebentrost, P.; Aspuru-Guzik, A. Communication:
Exciton-phonon information flow in the energy trans-
fer process of photosynthetic complexes. J. Chem. Phys.
2011, 134, 101103.
28 May, V.; Kuhn, O. Charge and energy transfer dynamics
in molecular systems; Wiley-VCH: Weinheim, 2004.
29 Mukamel, S. Principles of nonlinear optical spectroscopy;
Oxford University Press: New York, 1999.
30 Cho, M. Two-dimensional optical spectroscopy; CRC
Press: Boca Raton, 2009.
6
31 Hein, B.; Kreisbeck, C.; Kramer, T.; Rodr´ıguez, M. Mod-
elling of Oscillations in Two-Dimensional Echo-Spectra of
the Fenna-Matthews-Olson Complex. New J. Phys. 2012,
14, 023018.
32 Adolphs, J.; Muh, F.; Madjet, M. E.-A.; Renger, T. Calcu-
lation of pigment transition energies in the FMO protein.
Photosynth. Res. 2008, 95, 197.
C.;
33 Kreisbeck,
ics
Lab
HEOM).
http://nanohub.org/resources/16106.
for
2013;
Kramer,
T.
Light-Harvesting Complexes
doi:10.4231/D3RB6W248
Exciton Dynam-
(GPU-
at
online
34 Tiwari, V.; Peters, W. K.; Jonas, D. M. Electronic reso-
nance with anticorrelated pigment vibrations drives photo-
synthetic energy transfer outside the adiabatic framework.
Proc. Natl. Acad. Sci. 2012, doi:10.1073/pnas.1211157110.
35 Yuen-Zhou, J.; Krich, J. J.; Aspuru-Guzik, A. A witness for
coherent electronic vs vibronic-only oscillations in ultrafast
spectroscopy. J. Chem. Phys. 2012, 136, 234501.
36 Yuen-Zhou, J.; Aspuru-Guzik, A. Quantum process tomog-
raphy of excitonic dimers from two-dimensional electronic
spectroscopy. I. General theory and application to homod-
imers. J. Chem. Phys. 2011, 134, 134505.
37 Yuen-Zhou, J.; Krich, J. J.; Mohseni, M.; Aspuru-
Guzik, A. Quantum state and process tomography of en-
ergy transfer systems via ultrafast spectroscopy. Proceed-
ings of the National Academy of Sciences 2011, 108,
17615 -- 17620.
|
1802.01067 | 1 | 1802 | 2018-02-04T03:34:47 | Computation of a Theoretical Membrane Phase Diagram, and the Role of Phase in Lipid Raft-Mediated Protein Organization | [
"physics.bio-ph",
"q-bio.SC"
] | Lipid phase heterogeneity in the plasma membrane is thought to be crucial for many aspects of cell signaling, but the physical basis of participating membrane domains such as "lipid rafts" remains controversial. Here we consider a lattice model yielding a phase diagram that includes several states proposed to be relevant for the cell membrane, including microemulsion - which can be related to membrane curvature - and Ising critical behavior. Using a neural network-based machine learning approach, we compute the full phase diagram of this lattice model. We analyze selected regions of this phase diagram in the context of a signaling initiation event in mast cells: recruitment of the membrane-anchored tyrosine kinase Lyn to a cluster of transmembrane of IgE-Fc{\epsilon}RI receptors. We find that model membrane systems in microemulsion and Ising critical states can mediate roughly equal levels of kinase recruitment (binding energy ~ -0.6 kBT), whereas a membrane near a tricritical point can mediate much stronger kinase recruitment (-1.7 kBT). By comparing several models for lipid heterogeneity within a single theoretical framework, this work points to testable differences between existing models. We also suggest the tricritical point as a new possibility for the basis of membrane domains that facilitate preferential partitioning of signaling components. | physics.bio-ph | physics | Computation of a Theoretical Membrane Phase Diagram, and the Role of Phase in
Lipid Raft-Mediated Protein Organization
Eshan D. Mitra†1, Samuel C. Whitehead‡, David Holowka†, Barbara Baird†*, James P.
Sethna‡*
†Department of Chemistry and Chemical Biology, Cornell University, 122 Baker Laboratory,
Ithaca, NY, 14853
‡Department of Physics, Cornell University, 109 Clark Hall, Ithaca, NY, 14853
1Current address: Los Alamos National Laboratory, P.O. Box 1663, Los Alamos, NM 87545
*Corresponding authors:
B. Baird, [email protected]
J. P. Sethna, [email protected]
Abstract
Lipid phase heterogeneity in the plasma membrane is thought to be crucial for many aspects
of cell signaling, but the physical basis of participating membrane domains such as "lipid
rafts" remains controversial. Here we consider a lattice model yielding a phase diagram that
includes several states proposed to be relevant for the cell membrane, including
microemulsion – which can be related to membrane curvature – and Ising critical behavior.
Using a neural network-based machine learning approach, we compute the full phase
diagram of this lattice model. We analyze selected regions of this phase diagram in the
context of a signaling initiation event in mast cells: recruitment of the membrane-anchored
tyrosine kinase Lyn to a cluster of transmembrane of IgE-FcεRI receptors. We find that model
membrane systems in microemulsion and Ising critical states can mediate roughly equal
levels of kinase recruitment (binding energy ~ -0.6 kBT), whereas a membrane near a
tricritical point can mediate much stronger kinase recruitment (-1.7 kBT). By comparing
several models for lipid heterogeneity within a single theoretical framework, this work points
to testable differences between existing models. We also suggest the tricritical point as a new
possibility for the basis of membrane domains that facilitate preferential partitioning of
signaling components.
1
Introduction
The lateral organization of cell plasma membranes, which contributes crucially to their
functions, is regulated by membrane proteins and lipids as well as by attachment to the
cytoskeleton and by communication with membrane trafficking and other cellular processes.
A primary component of membrane organization appears to be the collective properties of the
lipid populations, and this has been examined experimentally and theoretically, as described
in numerous recent reviews (see 1,2 and reviews cited therein and elsewhere in this paper).
Whereas the diameter of a constituent lipid is about 1 nm, the bulk of experimental evidence
suggests that mammalian plasma membranes contain phase-based domains on the order of
10-200 nm in length.3–6 This heterogeneity has been related to studies of simpler model
membranes composed of a high melting point (Tm) lipid, a low Tm lipid and cholesterol,
considered to serve to as an approximation of plasma membrane lipids.7 Varying relative
amounts of these three types of lipids has yielded phase diagrams showing regions of
separation between phases characterized as liquid ordered (Lo, more high-Tm lipid and more
cholesterol) and liquid disordered (Ld, more low-Tm lipid).8–10
Studies on giant plasma membrane vesicles (GPMVs), which are isolated from cells,
exhibit micrometer-scale regions of Lo-like and Ld-like phase character.11 Similar
fluorescence microscopy studies on intact cells under physiological conditions do not detect
Lo/Ld separation above the diffraction limit, possibly due in part to their dispersal by
cytoskeletal attachment in cells.12 However, electron spin resonance (ESR) studies on intact
cells provide evidence for coexisting Lo and Ld domains.13 In cell plasma membranes these
nanometer-scale phase-like domains are thought to be coalesced or stabilized as a result of
an external stimulus (e.g. antigen cross-linking of immune receptors), and to play an essential
role in stimulated cell signaling, by facilitating colocalization of membrane proteins that
partition into the same Lo-like domain, and separating them from those that partition into Ld-
like domains.6,14 We are particularly interested in cases where induced interactions between
multiple Lo-preferring components stabilize these domains, thereby recruiting other Lo-
preferring components. Such lipid-mediated segregation has been implicated in many
mechanisms of membrane protein signaling, including immune receptors,15,16 G-protein
coupled receptors,17 the oncogenic GTPase Ras,18 and others. A generic term that has
emerged for plasma membrane domains of Lo-like character is "lipid rafts," and, although the
size, dynamics, and other features of these structures in functional cells surely vary
compared to those in model membranes, the lipid phase properties are expected to be
2
similar.
Theories of raft formation
Despite the centrality of lipid-based membranes to cell biology, there remains no
consensus on the physical basis of lipid domains. As described above, formation of lipid rafts
has been tied to the observation of phase separation in model plasma membranes, including
giant unilamellar vesicles (GUVs)8–10,19,20 and GPMVs11,21. In addition to the simplest forms of
two-phase coexistence, these systems exhibit a rich variety of phase behavior, including,
microemulsionsa,19,22 lamellar phases (also called modulated phases),23 and critical
phenomena.21 Moreover, despite recent advances in experimental techniques (for recent
reviews, see 16,24), lipid rafts in cell plasma membranes remain a difficult system to investigate
– the dynamics and complexity of real cell systems notwithstanding, the 10-200 nm
dimension of rafts4 prevents direct observation via conventional light microscopy. Thus, the
goal for a theoretical consideration of lipid raft physics should provide comparisons and
hypotheses that are amenable to testing with the currently available tools.
Towards this end, various theoretical models have been proposed to describe raft-like
phenomena. However, due to the lack of direct experimental data on lipid rafts, the set of
theories that are consistent with observation is relatively unconstrained – models that
disagree on the fundamental physics of raft formation can give qualitatively similar results
that agree with extant experimental work.1 One theoretical viewpoint is that lipid rafts are
mediated by membrane curvature,25–27 which makes the interface between immiscible
membrane domains more energetically stable. It has also been proposed that a surfactant
species could provide a similar interface between domains.28 Both of these viewpoints
suggest that rafts exist as part of a microemulsion phase, in which nanoscopic domains of a
characteristic size are stabilized due to the curvature or surfactant. An alternate hypothesis
a Some groups describe the presence of "nanodomains,"19 a state of two-phase coexistence consisting of
nanoscopic domains of a characteristic size, rather than a microemulsion, which is defined as a one-phase
state with domains of a characteristic size.
The difference in terminology arises from a difference in the definition of the location of the phase boundary.
Theoretical physicists commonly define a phase based on the average of some order parameter, which is
calculated over a long length scale. If this length scale is larger than the characteristic domain size, then the
domains are averaged out in this calculation, leading to the conclusion that the system consists of a single
phase, and the designation of "microemulsion". However, some experimental groups define a system to be
in two-phase coexistence whenever an experimental technique (e.g. FRET, which has a detection length
scale of ~2-8 nm) detects the presence of two components. 22 Analysis of the same "microemulsion" system
with small characteristic domains would indeed give detection of two distinct components, leading to the
conclusion of two-phase coexistence, and the label of "nanodomains."
In this study, we use the term "microemulsion," but note that the same area of the phase diagram could be
deemed "nanodomains" if one adopted an empirical definition of two-phase coexistence such as is used in
22.
3
suggests that rafts are formed from critical
fluctuations in membrane composition, a result of
proximity of the membrane to a 2D Ising critical
point.21,29 Experimental studies have provided
support for both an Ising critical point21 and
interactions consistent with a microemulsion.22,23
We define lipid rafts as nanoscale domains
concentrated with Lo-preferring components, and
as described above, these domains can serve to
colocalize membrane proteins that partition
similarly.
To analyze the functional consequence of
Figure 1. Signal initiation by IgE-FcεRI.
IgE-FcεRI are cross-linked by an external
antigen. The resulting cluster of receptors
stabilizes a lipid raft that enables the
recruitment of Lyn. Lyn performs the initial
phosphorylation steps that transmit the
signal to more downstream signaling
partners.
rafts in depth, we focus on the tractable example
of transmembrane signaling mediated by the IgE-
FcεRI receptor in mast cells. Physiologically, this
stimulated transmembrane coupling activates
cellular signaling pathways involved in allergic immune responses (reviewed in 30–32). The
mast cell is stimulated when specific multivalent ligands (antigen) physically cross-links
several IgE-FcεRI receptors together in a cluster. This clustering causes recruitment of the
kinase Lyn, which is anchored to the inner leaflet of the plasma membrane and when
recruited, phosphorylates the receptor, thereby activating downstream signaling events
(Figure 1). This kinase recruitment is thought to be raft-mediated: both the cross-linked
receptors and the kinase preferentially partition into Lo-like membrane domains, facilitating
their coupling on the plasma membrane.15,33,34 The mast cell system serves as an example of
a more general paradigm in cell biology, in which the orchestrated co-clustering of membrane
proteins due to an external stimulus leads to initiation of transmembrane signaling.18
Model
In this work, we address some of the ambiguities in the physics of lipid raft formation
by a comparative approach. We employ a lattice model originally described by Gompper and
Schick,35,36 which can be used for simultaneous evaluation of both microemulsions and critical
phenomena. In addition, this model captures features such as a lamellar (modulated) phase
and two-phase coexistence observed in other membrane studies. Moreover, the model
exhibits a tricritical point – defined as the termination of a three-phase coexistence regime in
4
a phase diagram – which we suggest has interesting implications for stimulated cell signaling.
The model consists of a two-dimensional square lattice with the Hamiltonian (Eq. 1
with variables defined below)
H=∑
(Hσ i+Δσi
2)+∑
i
i,j
(−Jσ i σ j−Kσ i
2 σ j
2)+∑
L σi(1−σ j
2)σk
(1)
i,j,k
Each site on the 2D lattice can take a spin value σ of -1 (black pixel), 0 (gray pixel), or
1 (white pixel). Black and white pixels represent membrane components favoring Ld and Lo
domains, respectively. Gray pixels represent a surfactant when surfacant strength L is greater
than 0, or simply a molecule with neutral domain preference when L = 0. The summation over
i is over all sites in the lattice; i,j is over all nearest neighbors; i,j,k is over all groups of three
adjacent pixels in a straight horizontal or vertical line. We equate one lattice unit to a length of
1 nm, the approximate diameter occupied by one membrane lipid molecule.
Each of the five parameters of this model – H, Δ, J, K, and L – has units of energy. We
consider only non-negative values for J, K, and L, while H and Δ can take any value. The
external fields H and Δ control the composition of the lattice. H controls the relative
abundance of σ=-1 (black pixels) and σ=1 (white pixels), while larger Δ increases the
concentration of σ=0 (gray pixels). We consider this model in the grand canonical ensemble:
our simulation box represents one section of the membrane, so it makes sense that the
number of each type of molecule can fluctuate, analogous to molecules diffusing in and out of
the box. The coupling J between adjacent pixels represents the usual Ising model coupling,
which, for a membrane model, is the preference for molecules that prefer Lo domains to be
adjacent to other molecules that prefer Lo domains (and similarly for Ld-preferring
molecules). J can also be thought of as equal to the line tension between black and white
pixels times a distance of 1 lattice unit (1 nm). K is a two-pixel interaction that gives a
favorable energy to adjacent non-gray pixels. For a particular concentration of gray pixels, a
higher value of K makes it more favorable to have those gray pixels adjacent to each other. L
controls the strength of gray pixels as a surfactant; this term contributes a nonzero value only
when a gray pixel (σ =0) sits between two non-gray pixels (σ =±1), and is favorable when the
two non-gray pixels have different signs. Thus, increasing L > 0 makes it more favorable for
gray to sit between black and white.
In our implementation, rather than choosing a value for J, we choose a value for
temperature T in units of J/kB, and J is set accordingly. The other parameters H, Δ, K and L
are chosen in units of J. Boltzmann's constant kB is set to unity.
5
Note that when L=0 Eq. 1 reduces to the Blume-Emery-Griffiths model.37 With K also
set to 0, Eq. 1 becomes the Blume-Capel model.38,39 With Δ = -∞, corresponding to no gray
pixels present, Eq. 1 reduces to the Ising model.
Phases in the Lattice Model
When the lattice model of Gompper and Schick was initially described, it was possible
to extract some key features of the phase diagram, most notably the location of the critical
line, by finite size scaling.35 With the great increase in the power of computational resources
since that time, it has become possible for us to address the model more globally by
simulation.
We further take advantage of neural networks, which have become a powerful
machine learning technique, leading to the development of computational tools to address
challenging problems such as image recognition.40 In image recognition, a neural network is
trained to read the pixel values of an image, and output a label corresponding to what the
image shows, such as distinguishing between a cat and a dog. Similarly, neural networks
have been trained on simulated snapshots of phase models in physics, to output a label
corresponding to which phase the snapshot represents. When this phase classification is
performed for snapshots at a large number of model parameter sets, one obtains the phase
diagram of the model.41,42 With this methodology, we label a region of parameter space as a
distinct "phase" if the neural network is able to distinguish simulation snapshots in that region
from snapshots representing other phases. This definition is not always equivalent to a
thermodynamic definition of a phase (i.e. based on the value of order parameters), but rather
puts a greater emphasis on visually identifiable, qualitative differences in system properties.
Based on our neural network analysis, we describe eight phases (distinguishable
qualitative behaviors specified below) that the model (Eq. 1) produces. We name these as
follows: Within the fluid phase, all three components are well-mixed, with only short-range
interactions between them. The black phase and white phase consist of nearly all black pixels
and white pixels, respectively. When H=0, the Hamiltonian (Eq. 1) is symmetric with respect
to exchanging black and white, and so these phases are seen in a state of two-phase
coexistence. The gray phase consists of nearly all gray pixels. The microemulsion "phase"
consists of black and white domains stabilized by a boundary of surfactant. The critical
"phase" consists of fluctuating black and white domains, resulting from close proximity above
a critical phase transition. Note that the microemulsion and critical "phases" are not
6
thermodynamically distinct from one another or from the fluid phase.b 7,43 However, because
we are interested in the qualitative nature of domains that could be relevant for membranes,
we choose to consider them separately. The lamellar phase, is similar to a microemulsion
"phase" in that the surfactant separates the black and white domains, but instead of
enclosed, roughly round domains, the two domains exist as long stripes. Finally, the crystal
phase includes the behavior in which rectangular domains of black and white exist, separated
by a meshwork of surfactant.
Applying the model to cell signaling
Here, we apply the methodology of neural networks to the Gompper and Schick lattice
model, with the ultimate goal of understanding how different qualitative phase behaviors in
membranes compare in their capacities to mediate cell signaling through membrane
receptors (Figures 1 and 2c). The neural-network-derived phase diagram labels regions of
parameter space according to their distinctive behaviors, as described in the previous
section. We use this diagram to focus on sections of parameter space that are proposed to
be relevant for plasma membrane heterogeneity, in particular the microemulsion and critical
"phases". At these interesting points, we perform Monte Carlo simulations to calculate the
energy associated with recruitment of an inner-membrane-anchored kinase (Lo-preferring)
into a transmembrane receptor cluster (also Lo-preferring), as in the mast cell signaling
system. Note that these recruitment energies-in contrast to binding energies associated with
chemical bonds-are associated with long range forces: Proteins are recruited into an
energetically favorable region, without orienting and binding directly to specific sites on
proteins that stabilized the energetically favorable region. Also the energies we calculate are
non-specific – Lo and Ld preferring proteins will share the same interactions as a group, and
their structure details would only determine the degree of preference. Thus these long range
forces allow nonspecific interactions that are restricted only in terms of the components'
phase preference, as for co-localization in lipid rafts.
Monte Carlo methods allow us to explore the protein energetics semi-quantitatively
throughout the phase diagram. Moreover, the recruitment energies that we calculate agree
with exact conformal field theory results near the Ising critical point,44 and hence should
quantitatively describe experimental systems near critical points. And while our simulations
focus on a simplified model of clustered receptors, near critical points our results are
b Those who define microemulsion-like states as two-phase nanodomains (see footnote a), would instead say
that the microemulsion and critical "phases" are part of the two-phase coexistence between the black and
white phases.
7
universal, and are thus generalizable to a broad range of phenomena associated with
membrane heterogeneity. In total, this method of recruitment energy calculation allows us to
evaluate how the qualitative behavior of the plasma membrane relates to its capacity to form
lipid rafts that can be stabilized (e.g., by clustered receptors) to mediate biologically relevant
signaling.
The neural network approach is uniquely suited for this goal, offering a number of
advantages over more traditional analysis approaches. First, it is capable of exploring large
areas of parameter space at low computational cost. Second, it is able to detect qualitative
changes in model behavior, such as microemulsions, even if those changes do not
correspond to a true thermodynamic phase transition. These qualitative differences have
important consequences for cell signaling that is facilitated by membrane organization.
Methods
Monte Carlo Simulations
Snapshots of the lattice model35 were generated by the Metropolis algorithm. The
length of the simulation was counted in sweeps, where, in each sweep, each lattice site has
on average 100 opportunities to be flipped (total of 90000 individual proposed moves for a
30x30 lattice,c Figure 2a). Each proposed move consisted of randomly choosing a lattice site
and a target value (one of {-1, 0, 1} that was not the current value at the site). The move was
performed with probability min(1, e-ΔU/T) where ΔU is the change in the Hamiltonian energy
(Eq. 1) resulting from the move.
To generate a single independent snapshot, the lattice was randomized, then 100
sweeps were run to equilibrate, and the final result was saved. To generate correlated
snapshots, additional sweeps were run after equilibration, and a sample was saved after
each sweep. Such snapshots are correlated because a single sweep is not enough to fully
reequilibrate the lattice.
Neural Network Training
We chose the cross-section H/J=0, K/J=2, L/J=3 (see Hamiltonian, Eq. 1) for training
c The lattices we use for mapping phase diagrams are small; the size was chosen to capture the correlations
on length scales of interest to protein aggregation, and for convenient training of the network. Phases
without structure on long length scales should be well described by our small simulations; we would expect
shifts in boundaries of microemulsion phases, for example, only when the modulation approaches 30 pixels.
Near critical points all length scales are important for the physics, but we show that the phase boundaries
converge fairly rapidly. The shift in the effective critical temperature in a system of size L goes as L1/ν, so for
the Ising critical point with ν=1 we expect 3% shifts in phase boundaries for a 30x30 system (beyond the
precision of our methods), and near the tricritical point with ν=5/9 we find even smaller shifts.
8
because this is close to the cross-section described by Gompper and Schick35 as containing
examples of all major phases of the model. Generation of the neural network training data
was an iterative and somewhat heuristic process. We started by sparsely sampling a large
region of (Δ,T) space in the H/J=0, K/J=2, L/J=3 plane and labeling phases manually, to get a
general sense of the layout of the phase diagram. This allowed us to find regions where we
were highly confident about the correct classification, and we used these regions for training
data. In the case of the microemulsion phase, this included checking that the correlation
function had a local minimum.d After the first round of training and testing, we examined
snapshots from different points in the phase diagram to visualize where errors occurred, and
we added further training data at appropriate points to reduce these errors. For example, we
initially did not include the crystal phase consisting of black and white rectangles, as this
phase was not described in previous work. We identified this as a separate phase after it was
labeled as fluid phase in earlier tests. The final training data set is shown in Figure S1a,
overlaid on the final phase diagram. At each chosen set of training parameters (156 sets in
total), 100 independent samples were acquired for training, for a total of 15600 samples in
the training set.
Note that, despite the heuristic approach to generating the training data, it is not the
case that we could generate an arbitrary different phase diagram simply by changing the
training data. Rather, the phase diagram reflects real, qualitative differences in the behavior
of the system. In our experience, training with a bad training set (e.g., containing different
phases labeled as the same phase) leads to an obviously bad phase diagram, in which some
regions contain different adjacent pixels classified as different phases with low confidence
(quantified as described below).
Two types of training data were acquired for use in training two separate networks. In
one data set (the snapshot approach, phase diagram shown in Figure S1c), simply 100
independent snapshots per parameter set were saved. In a second data set (the averaged
approach, phase diagram shown in Figure S1b), 100 independent groups of 10 correlated
snapshots each (as described in Monte Carlo Simulations, above) were acquired. The 10
snapshots were averaged to give one average image for the data set. Broadly speaking, this
averaging has the effect of smoothing out random fluctuations, allowing the network training
to focus on more constant aspects of each phase.
d The appearance of this oscillation in the correlation function is one (admittedly somewhat arbitrary) definition of a
microemulsion suggested by Gommper and Schick.32
9
Figure 2. Schematic of the methodology used in
this study. (a) Schematic of the neural network
(NN) used for phase prediction. The pixel values
from a Monte Carlo simulation on a 30x30 lattice
serve as inputs. (Black, white, and gray pixels
are rescaled to non-negative values for these
simulations as described in Methods). The
network is trained using 2 hidden layers of 100
nodes each. The network contains 6 outputs,
corresponding to its confidence that the input
represents each of the 6 possible phases. Each
pixel in the simulation box has dimensions of 1
nm x 1 nm. (b) At each point in parameter space
(square pixels), the neural network was run on
Monte Carlo simulation results to label the phase
(c) Schematic of simulations used to calculate
kinase binding energy by Bennett's method. The
simulated system consists of two separate
boxes, one representing the membrane near the
receptor cluster (left), and another representing
a section of the membrane at infinite distance
(right). The teal and magenta proteins' σ values
are fixed white, while the rest of the lattices are
Monte Carlo sampled. We use Bennett's method
to calculate the free energy difference between
State 1 (kinase at infinity) and State 2 (kinase
inside cluster). The dimension of each pixel in
the simulation boxes is 1 nm x 1 nm. The solid
lines linking panel a to panel b and panel b to
panel c show one example of how a phase is
determined and used in Bennett's calculation.
The neural network code used is the
implementation of 45, also available online at
https://github.com/mnielsen/neural-networks-
and-deep-learning . Each training sample was
converted into an input vector of length 900
containing the values at each site of the 30x30
lattice, and a target output vector of length 6,
consisting of 1 at the index of the correct phase,
and 0 for all other values. The values of the input
vector were rescaled such that black = 0, gray =
0.5, white = 1, in order to provide all non-
10
negative inputs to the network (Figure 2a). The feed-forward neural network contained two
hidden layers of size 100 each, made up of sigmoid neurons. We performed 25 epochs of
training. In each epoch, the training data were randomly divided into mini batches of size 10.
With each mini batch, stochastic gradient descent was performed by a backpropagation
algorithm with a learning rate η = 0.06. We use a cross-entropy cost function, with an L2
regularization parameter of λ = 0.04 to avoid overfitting. To avoid stopping the stochastic
training at a bad point, if the final classification accuracy was worse than 0.85, extra epochs
were run, one at a time, until 0.85 was reached. For the snapshot approach, we instead used
a threshold of 0.9. This method resulted in at most 5 (typically 0-2) extra epochs added. 10
instances of the neural network were trained independently on the same training data set.
When working with the test data, we took the average output of the 10 instances.
Neural Network Phase Diagram Generation
Test data were generated by the same Metropolis method as the training data. At each
point in parameter space (H, K, L, Δ, T; Eq. 1) where we sought to determine the phase, 5
snapshots or correlated averages were generated. These were fed as input into the neural
networks, yielding output vectors with 6 elements in the range [0,1]. In these output vectors, a
higher value at a particular index indicates that the point more likely belongs to the
corresponding phase. Output vectors were averaged over the 5 samples and 10 network
instances to arrive at a single final output vector (Figure 2a). The point was classified as the
phase corresponding to the maximum value in the output vector (Figure 2b). The
classification confidence was calculated as the maximum value in the output vector, divided
by the sum of the output vector. When rendering the phase diagrams, the phase classification
determined the color – red, green, blue, orange, pink, or yellow. The RGB value of the base
color was multiplied by the classification confidence, such that a brightere color represents
more confident classification. For example, a point classified as lamellar (red, RGB = (0.8,
0.4, 0.0)) with confidence 0.8 would be rendered as RGB = (0.64, 0.32, 0.0).
The averaged approach was more effective than the snapshot approach. With the
snapshot approach, we could only distinguish 4 phases: Fluid, Black/White, Gray, and a
single region covering Lamellar, Microemulsion, and Critical (Figure S1c). With the averaged
approach, we could distinguish six phases (Figure S1b), but we had low confidence in the
distinction between the Fluid and Gray phases (Figure S1d). To combine these, on testing
e Note that we use the term brightness here in the sense of the HSB (Hue, Saturation, Brightness) representation of
colors. HSB and HSV (Hue, Saturation, Value) are equivalent representations, so scaling the brightness is synonymous
with scaling the value.
11
data, we used the Gray output from the snapshot approach, and the other 5 outputs from the
averaged approach. This gave the final phase diagram that we believe most completely
describes our understanding of it after our work with both these approaches.
Binding Energy Computation
We consider the binding energy to be the difference in free energy between a single
white pixel (spin +1) with a set cluster of three other white pixels, compared to that single
white pixel being at an infinite distance from that set cluster (Figure 2c). We call the set
cluster "receptors" and the designated single pixel, "kinase." To compute this binding energy
by Bennett's method,46 simulations were performed on the four separate lattices shown in
Figure 2c: State 1 consists of a 50x50 lattice containing the set cluster of receptors, and a
separate 30x30 lattice containing the kinase. State 2 consists of a 50x50 lattice containing
the kinase within the cluster of receptors, and a 30x30 lattice empty of the kinase. Note, for
the 30x30 boxes (Figure 2c, right), the smaller lattice size was permissible because these
boxes only ever contain one designated white pixel, which affects the lattice on a shorter
length scale than the full receptor cluster. Samples were generated by the Metropolis
algorithm in the same way as the neural network training data, but the predefined receptor
and kinase proteins were required to remain white. Any proposed move that attempted to flip
one of these spins was automatically rejected.
The free energy ΔF, corresponding to the binding energy, is computed according to
the following formula.
e−(ΔF−C )/(k BT )=
⟨f (( ΔU 1→2−C)/(kB T))⟩1
⟨f ((ΔU 2 →1+C)/(k BT ))⟩2
(2)
Here, C can be any constant, with the fastest convergence achieved when C ≈ ΔF. We
choose C = -0.5 kBT, and choose f as the Fermi-Dirac function, f ( x)=1/(1+ex) as suggested in
46. The numerator is calculated as an ensemble average from simulations of state 1 (Figure
2c, top). ΔU1→2 for each sample is the energy change associated with exchanging the kinase
and a pixel at the center of the cluster (corresponding to the kinase position in state 2).
Likewise, the denominator is calculated from simulations of state 2, and ΔU2→1 is the energy
change associated with exchanging the kinase located within the cluster and the pixel
corresponding to its position in state 1.
Note that the two separate boxes that make up each state in Figure 2c can be
generated independently, and we use this to our advantage. We initially generated the same
12
number of samples of the 50x50 box and the 30x30 box. Then each 50x50 box was paired
with 10 different 30x30 boxes, increasing the number of samples of the state by a factor of
10. These samples are not independent, but they still follow the correct Monte Carlo statistics.
For calculating binding energy at each parameter set to be tested, simulations were
performed for 5000 sweeps, a sample was saved every sweep, and the lattice was reshuffled
every 10 sweeps. After data expansion, this gave 50000 non-independent samples of each
state, to be used in the Bennett calculation.
Results
Neural Network Phase Identification
We trained neural networks to classify the output of a Monte Carlo simulation of the
Gompper and Schick lattice model,35 according to the phase that the simulation represents. A
schematic of the network and an example of a resulting phase diagram are shown in Figure
2a, b. The σ values from a 30 by 30-pixel Monte Carlo snapshot (generated by the standard
Metropolis method)47 were used as 900 inputs to the network. Training data consisted of
15600 such snapshots, which represented typical examples of each phase of interest (Figure
S1). The network was trained with 2 hidden layers of 100 nodes each, and an output layer of
6 nodes, corresponding to the six phases of interest in the phase diagram. Alternatively,
instead of single Monte Carlo snapshots, we used input consisting of the average of 10
correlated snapshots from consecutive simulation steps. This method tended to be more
accurate in most cases, and our final reported phase diagrams make use of some output
from both types of networks. Our procedures, including training of the neural networks, are
further described in Methods.
We initially evaluated the lattice model with H/J = 0, K/J = 2, L/J = 3, ranging over T/J
and Δ/J values of order 1. In the original description of the model,35 this cross-section was
found to contain examples of all phases present in the model.
Our neural network was able to confidently label six distinct regions of the phase
diagram (Figure 3), corresponding to the eight phases described in the Introduction: fluid,
lamellar, gray, crystal, black / white, and microemulsion / critical. The network was not able to
determine a distinct boundary between microemulsion and critical fluctuations, so the single
microemulsion / critical label was applied to both. At larger values of Δ, the region is a
microemulsion, while, at smaller values of Δ, the system shows fluctuating domains
consistent with close proximity to an Ising critical point. Instead of a clear boundary between
13
Figure 3. Phase diagram of the lattice model. The color of each pixel with specified (Δ/J,
T/J) coordinates indicates the phase at that point, as determined by the neural network.
Pixels with a higher brightness indicate a higher level of confidence in the classification.
Snapshots show typical examples of each of the phases, corresponding to the black
points on the phase diagram.
14
the critical fluctuation and microemulsion behaviors, the regions blend into one another
smoothly. Because two models often used to explain lipid rafts – microemulsion and critical
phenomena1 – are included within this phase, it is highly relevant for membrane-related
questions.
The network applied the same black / white label to both the black phase and white
phase. Because the training data contained examples of two-phase coexistence, including
snapshots of both black phase and white phase with the same classification, the neural
network was trained to apply the same label to both. With H=0, the black / white classification
represents two-phase coexistence between the black and white phase, while with H>0, the
white phase does not exist, and the label represents only the black phase (conversely for
H<0). Finally, we note that the network applied the crystal label to the limit of the lamellar
phase in which the components alternate with period of one lattice unit.
Exploring the Phase Diagram
We used our neural network to compute other cross-sections of the phase diagram
and thereby gain a more complete perspective on the entire parameter space. Remarkably, it
was not necessary to retrain the network to work with these other cross-sections. We found
that the original network trained at H/J=0, K/J=2, L/J=3 accurately identifies the phases
elsewhere in the phase diagram, for all H, K, and L values considered in this study.
Varying the surfactant strength L changes the topology of the phase diagram (Figure
4). At zero or low L (L/J = 1.5), the lamellar phase does not exist, and the black/white phase
directly borders the gray phase. At zero L, a tricritical point exists at the intersection of the
fluid, black/white and gray phases. At higher L (L/J = 3), we reach the case shown in Figure
3, in which the lamellar and crystal phases exist between the black/white and gray phases. At
even higher L (L/J = 6), the system becomes a crystal for nearly all values of Δ and T tested,
maximizing the number of surfactant interactions.
With K=0 and L=0, the model reduces to the more widely studied Blume-Capel model
(Figure 5), in which gray pixels are neutral in their interactions with white and black pixels. In
our diagram, the region between black/white coexistence and the fluid phase can be
identified as a critical transition by virtue of the yellow critical region appearing between the
blue and orange regions. Note that microemulsions are not possible with L=0, and therefore
the entire yellow region in this cross-section represents Ising critical behavior. The critical line
occurs at the boundary between the blue and yellow regions in Figure 5a. With H/J = 0.1, the
critical "phase" disappears, correctly showing that at L=0, H>0, there is no longer a critical
15
Figure 4. Cross-sections of the phase diagram at
varying values of the surfactant strength.
Surfactant strength L/J is varied 0 to 6, with
constant H/J = 0, K/J = 2. Colors have the same
meaning as in Figure 3.
phase transition (Figure 5b).
When L is increased with K=0 (Figure S2),
the phase diagram has topology similar to as the
case with K/J = 2, although the phase boundaries
occur at lower Δ. Finally, we considered some
additional cross-sections at positive H (Figure S3).
We note that with H/J=0.1, K/J=2, L/J=3, some
yellow region remains at high Δ. Presumably this
indicates microemulsion behavior, because a critical
line is not expected to exist at nonzero H. At higher
H (H/J=0.5), the black/white classification (here
representing only the black phase) grows to
encompass most of the parameter space examined
in this range of Δ and T.
Quantifying Protein Recruitment in Terms of
Preferential Partitioning
Having calculated the phase diagram for the
lattice model, we turned to our questions related to
biological function. In particular, we compare the
effectiveness of lipid raft-mediated protein
reorganization at various points on the phase
diagram. As a specific test, we consider the case of
three receptors (such as IgE-FcεRI) cross-linked to
form a cluster; these are activated to initiate
transmembrane signaling only after recruiting a
membrane-anchored kinase (such as Lyn; see
Figure 1). We assume both the receptors and the
kinase prefer Lo-rich domains (i.e., lipid rafts), and
16
correspondingly, we represent them with white
pixels, which we place at selected, fixed positions in
the lattice. Lyn, is represented by one white pixel,
whereas each of the three receptors is represented
by 12 white pixels, corresponding to their relatively
larger size (Figure 2c). We calculate the binding
energy as the free energy associated with moving
the kinase into the middle of the three-receptor
cluster. A larger magnitude negative value indicates
a stronger contribution of lipid rafts to protein
colocalization at a particular point in the phase
diagram.
Similar to what we and others have done
Figure 5. Cross-sections of the phase
diagram for the Blume-Capel model
(K=0, L=0), with H/J = 0 or 0.1. Colors
have the same meaning as in Figure 3.
previously,44 we use Bennett's method46 (Eq. 2) to
calculate the free energy change. We do so here in
a more computationally efficient method than in
previous studies. In previous work, we calculated
the energy change stepwise, moving the kinase out
of the cluster, one lattice unit at a time, and
generating a profile of energy versus position in the
process.44 Here, we instead calculate the entire
energy in one step. Our simulated system (Figure 2c) consists of two separate boxes, one
containing the receptor cluster (left), and the other representing membrane at infinite distance
from the cluster (right). By Bennett's method, we compute the free energy to move the kinase
from the box at infinite distance (state 1) to the center of the cluster (state 2).
We used the phase diagram to assist in choosing points for Bennett simulations – we
ran a simulation at each point marked with a diamond in Figure 6. We focused our
simulations primarily on the microemulsion/critical region of the phase diagram, and, for
comparison, we performed simulations at a smaller number of points elsewhere in the phase
diagram. We additionally performed simulations in which a single white pixel was set (instead
of the receptor cluster) and calculated binding energies for a second white pixel to come into
proximity. We found these binding energies to be qualitatively similar but weaker compared to
the case with the cluster (Figure S4).
17
Our results for a kinase associating with a receptor cluster (Figure 2c) are shown by
the colors of the diamonds in Figure 6. In the Blume-Capel (H = K = L = 0) phase diagram
(Figure 6a), no microemulsion exists and we find roughly the same binding energy of ~ -0.6
kBT at all points along the Ising critical line at the boundary between the blue and yellow
regions. This corresponds to a modest increase in kinase concentration, by a factor of
e0.6≈1.8 . Along this critical line, the binding energy does not show a dramatic difference
above, versus below, the transition temperature (columns of diamonds along blue-yellow
boundary). Strikingly, as the tricritical point is approached (the box in Fig. 6a), we find a
dramatic, nearly threefold increase in the magnitude of the binding energy. The minimum free
energy of -1.7 kBT, is achieved at the tricritical Δ (1.9655 J) and 1.04 times the tricritical
temperature (0.634 J/kB). The corresponding increase in kinase concentration by a factor of
e1.7≈5.5 is much more significant than the 1.8 factor at an ordinary critical point. We suspect
that the distance of the optimum above the tricritical temperature, 1.04x, is a finite size effect,
as this value increases if the simulation box is made smaller. The true optimum might occur at
exactly the tricritical temperature (0.610 J/kB).
To validate our new application of Bennett's method (Eq. 2, Figure 2c), we also
calculated the energy profile at the tricritical point stepwise by Jarsynski's method,48 identical
to the method used in 44 (Figure S5). Due to the larger simulation box used in this method,
finite size effects are less of a concern. We found a binding energy of ~ -1.5 kBT with
Jarzynski's method, comparable to our result at 1.02x the tricritical temperature with
Bennett's method (Figure 6b). However, at the tricritical temperature, our application of
Bennett's method gives a binding energy of only -1.0 kBT, presumably due to finite size
effects at this temperature.
We compare these results to the first-order phase transition that occurs at H > 0
(Figure 6c), which yields a higher concentration of black (Ld-preferring) pixels than white (Lo-
preferring) pixels in the lattice. We found a similar binding energy of ~ -0.6 kBT above the
transition temperature in the fluid phase. However, we see a substantially stronger binding
energy as low as ~ -1.4 kBT upon entering the phase-separated state. In the context of
membranes, this would correspond to a situation in which most lipids on the membrane favor
the Ld phase, but our receptor/kinase proteins of interest favor Lo.
Finally, in Figure 6d, we consider the binding energy around the microemulsion/critical
region using the parameters of Figure 3 (H/J=0.0, K/J=2.0, L/J=3.0). To aid in the distinction
between microemulsion and critical "phase" in this cross-section, at selected points (marked
18
Figure 6. Kinase binding energy at selected points in the phase diagram. Each colored
diamond indicates the free energy change associated with moving a kinase into a cluster
of receptors (as in Figure 2c) at that point (Δ/J, T/J) in the phase diagram. The phase
diagram colors are rendered paler than in other figures to make the diamonds more clearly
visible. (a) Binding energy in the Blume-Capel model (K=0, L=0) for the cross-section H=0.
(b) Inset of (a) in the region around the tricritical point (black box in (a) ). The indicated
point (black arrow) with the minimum free energy of -1.7 kBT occurs at the tricritical Δ and
1.04 times the tricritical temperature. (c) Binding energy in the Blume-Capel model for the
cross-section in which the external field H/J = 0.1 favors black pixels, opposite to the
kinase and receptor preference for white pixels. (d) Binding energy in a cross-section that
includes microemulsions and lamellar phases (K/J = 2, L/J = 3). Within the
microemulsion/critical phase, marked points (magenta boxes) were analyzed with
correlation functions and visual inspection of simulation snapshots (Figure S6). Point A is
part of the critical "phase," point B is a microemulsion with length scale ~ 10, and point C
is a microemulsion with length scale ~ 4. At certain points in this cross-section (blue color
scale), including point C, the positive binding energy indicates that it is energetically
unfavorable to bring the kinase into the cluster.
19
A, B, and C in Figure 6d), we performed correlation function analysis (Figure S6). We
confirmed that point A is in the Ising critical region, and B and C are in the microemulsion
region. In the Ising critical region (i.e. the yellow region at low values of Δ, including point A),
we again find a binding energy of ~ -0.6 kBT, the same as the case with no surfactant strength
L (Figure 6a). As we move to higher Δ, corresponding to a microemulsion region, we find a
striking change. At a subset of the points in the microemulsion region (including point C), the
binding energy at becomes much weaker, even turning positive (unfavorable). Intuitively, this
happens when the characteristic length scale of the microemulsion is smaller than the size of
the set receptor cluster. Considering microemulsions with a longer length scale (near blue-
yellow boundary at Δ/J between 4 and 5, including point B), we find a binding energy of -0.6
kBT, comparable to that at an Ising critical point at lower values of Δ. Thus, the results
indicate that the binding energy associated with microemulsion behavior depends on how the
characteristic length scale of the microemulsion compares to the spacing of the clustered
receptors. It is also possible that a microemulsion exists at a length scale larger than our
30x30 nm snapshot used to generate the phase diagram. This would likely appear as phase-
separated in our diagram, and indeed would look equivalent to phase-separated from the
perspective of a cluster of size less than 30 nm. Based on our simulated results at points in
the phase-separated region, this case would also likely yield a value of around -0.6 kBT
(Figure 6d).
The Figure 6d cross-section contains no point comparable in binding energy to the
tricritical point in Figure 6a. The minimum binding energy achieved in this cross-section
(except perhaps at biologically irrelevant points at very low temperature) is ~ -0.6 kBT, which
occurs along the entire boundary between the black/white and microemulsion/critical regions.
This remains true if we more densely sample the entire length of the phase boundaries (data
not shown). Among all of the phase states tested, the tricritical point at H = K = L = 0 (Figure
6a) leads to the strongest possible binding energy for kinase and clustered receptors.
Discussion
Comparison to Published Results
We have generated the phase diagram for the Gompper and Schick lattice model
using relatively new neural network methodologies. It is important to consider how this
method compares to other more established methods for phase diagram determination. We
examine certain special cases of the model that allow for direct comparison of our phase
20
diagram to published phase diagrams obtained by other methods.
By taking Δ to -∞ (no gray pixels), and H=0, we have the Ising model, with the well-
known critical transition temperature of 2 / log(1 + sqrt(2)) ≈ 2.269 J/kB. Applying our existing
network to this case, we see the phase transition at close to the correct temperature (Figure
S7a). The network's confidence level for the Ising model is worse than optimal because this
network was trained to perform a more complicated classification on 6 phases, instead of the
2 phases (fluid and black/white coexistence) relevant to the Ising model. A different approach
is to train a network, solely on Ising model examples, to classify between only the fluid phase
and black/white coexistence. With this model, we distinguish the phases with high
confidence, and we nearly perfectly identify the transition temperature (Figure S7b). This level
of accuracy is comparable to previous neural network work on the Ising critical transition.41
The result for the Blume-Capel model (K=0, L=0; Figure 5) with H=0 is comparable to
results with this model from other methods. We find good quantitative agreement on the
location of phase boundaries with Beale's phase diagram from finite size scaling49 (Figure
S8). We also show the mean field theory solution37 for comparison. The tricritical point has an
upper critical dimension of three, meaning that mean field theory is expected to be inaccurate
near the tricritical point in this two-dimensional model.50 However, our calculated result is
much closer to the more accurate finite size scaling solution.
Our diagrams can also be compared to those obtained in Gompper and Schick's
original description of the model35 (Figure S9). Note that, to make this comparison it was
necessary to add the parameter K2, the equivalent of K between second nearest neighbors in
a straight line. This had no effect on the overall shape of the phase diagram, but shifted the
phase boundaries slightly. We find very good agreement on the location of the critical line in
all cross-sections with Gompper and Schick's transfer matrix approach. The original phase
diagram included a Lifshitz line, which the authors defined as the separation between Ising
and microemulsion regions. This helps us better interpret the combined microemulsion/critical
region in our phase diagram, which is in fact a microemulsion to the right of the Lifshitz line.
In other aspects of the phase diagram, the neural network approach provided new
information, and it revealed shortcomings of the original phase diagram. We note our new
placement of the lamellar phase (red) is qualitatively different from the Gompper and Schick
diagram, including a lobe that sits below the phase-separated state on the temperature axis.
We give a new boundary between the gray phase (pink) and the fluid phase (orange). Our
identification of the rectangular crystal phase (green) is entirely new, not addressed in the
21
original study (the diagonal crystal that our network labeled as part of this phase arguably
belongs in the lamellar phase, but the rectangular features are clearly a distinct phase).
Some of these novel features are relevant to the biological system of interest, while
others are not (such as the rectangular crystal phase, which likely exists only due to the use
of a square lattice), but all point to the strengths of global computational approaches in phase
diagram prediction, which allow direct comparisons. Theoretical techniques like finite size
scaling frequently focus on specific interesting areas of the phase diagram, such as the
critical line. In our neural network approach, we instead indiscriminately analyzed entire slices
of the phase diagram, extracting features in both critical and non-critical regions. This is
especially valuable for a problem such as biological lipid-based membranes, for which
different groups have proposed that the most relevant states are either near a critical point21,29
or away from a critical point.7,26,28
Finally, to further validate the application of this model to the study of lipid membranes,
we compare our neural-network-derived phase diagrams with the numerical and mean-field
phase diagrams produced in previous studies on the formation of lipid rafts (nanoscale
domains concentrated with Lo-preferring components, as defined in the Introduction). We
consider first microemulsion-based models, which propose that either surfactant-like lipid
species28 or membrane curvature7,26,27 stabilize the interface between different phase
domains. Importantly, the generality of our neural network approach means that we could in
principle explicitly reproduce the results of the different membrane models described above.
It should even be possible to train a neural network with multiple models simultaneously, a
potential avenue for future work. Here, however, we are interested in comparing the results of
our single-Hamiltonian neural network approach with results in the membrane modeling
literature.
How much agreement should we expect between the neural network trained on our
Hamiltonian (Eq. 1) and models with different Hamiltonian forms and explicitly different
energetic terms (e.g. composition-curvature interactions)? Due to the presence of gray pixels
as surfactants, our model most closely resembles models that make use of hybrid lipids,28,51
so we can ask how our model compares to curvature-based models, which are seemingly the
most different. As discussed above (Results), our Hamiltonian captures much of the physics
of other membrane models, including 2D Ising critical and tricritical behavior. In these critical
regimes, our Hamiltonian is equivalent to all others due to the universality of critical
behavior.50
22
Outside these critical regimes, in the biologically-relevant microemulsion phase, we
also expect qualitative agreement between our model and curvature-based models.
Intuitively, in a microemulsion regime, the gray pixels in our model will act analogously to
regions of curvature mismatch: in a system with droplets of one phase suspended in a
backdrop of another phase, the boundaries of the droplets will be regions of concentrated
surfactant-like interaction. In our model, this looks like a domain of either black or white pixels
encircled by a strip of gray pixels; in the curvature model, the picture is the same, except that
the gray pixels are replaced by a region of curvature change (this can be pictured as the
droplet "popping out" of the membrane). Importantly, in this regime we have a defined length
scale in both models: in ours, it arises from the concentration of gray pixels, while in the
curvature model it arises from the mechanical properties of the membrane.
Ultimately, however, the comparison of predicted phase behavior serves as the best
indicator of model similarity, and we find good agreement between the phase diagrams in the
literature (both curvature- and hybrid-lipid-based) and those generated by our neural network
approach (Figure 7). Our phase diagrams reproduce all the features found in these other
model frameworks, including Ising critical transitions, lamellar phases, two-phase coexistence
and tricritical phenomena. Moreover, the general topology of the phase diagrams is
consistent regardless of model choice-for instance, all models considered here predict a
lamellar phase separated from a microemulsion phase by an Ising critical line, with the
microemulsion phase, in turn, separated from an ordinary fluid phase by a boundary that is
not a true thermodynamic phase transition. This consistency with previously calculated phase
diagrams7,51 speaks to the generality of our approach, which allows us to describe and
compare a wide variety of membrane phenomenologies using a single model framework.
Application to Lipid Rafts
We set out with this model to analyze competing hypotheses on the physical basis for
formation of lipid rafts: does stabilization of nanoscale Lo-like domains arise from proximity to
a critical phase transition, or from nanodomains of a characteristic size, as in a
microemulsion? We found that in some ways, the two hypotheses are much alike. As
described in the Introduction, considerable evidence supports the view that lipid rafts serve to
recruit proteins to the correct place on the cell membrane, such as our example of Lyn kinase
recruitment into a set IgE-FcεRI cluster, where both components are Lo-preferring. Our phase
diagram shows that critical and microemulsion phase states can be equally beneficial
thermodynamically for this membrane purpose. As we showed, both can give about the same
23
Figure 7. Schematic comparison of our phase
diagram with those from the microemulsion literature.
(a) Mean-field phase diagram from a model with
hybrid lipid species acting as a surfactant. Adapted
from 51 (b) Mean-field phase diagram from a model
with curvature coupled to membrane composition to
produce a surfactant-like interaction. Adapted from 7.
(c) Phase diagram generated by our neural network
approach. X coordinate gives strength of surfactant-
like interaction in (b), or concentration of surfactant
species in (a) and (c). Y coordinate represents
temperature. Note that this is a schematic
representation, so the actual axes from the source
papers differ in scale and representation. For the
sake of comparison to the other models, we use
yellow here to represent only microemulsions, not
Ising critical behavior. The yellow-orange gradient in
(c) is used to schematize the ambiguity between
microemulsion and ordinary fluid phase, and
represents our best interpretation of the location of
the microemulsion state, taking into account the
neural network output (Figure 3), snapshots within
the phase diagram (Figure 3, yellow-bordered
panels), and the location of the Lifshitz line from
35 shown in Figure S9.
optimal binding energy of -0.6 kBT. We also showed
it is possible to sit in a region between microemulsion
and critical point with a classification that is
subjective. Gompper and Schick chose the Lifshitz
line as an arbitrary distinction for what qualifies as a
microemulsion, while our neural network was unable
to draw a sharp line between the two behaviors.
Our energy calculations make a clear prediction for
a difference between clearly critical and clearly
microemulsion states (at lower and higher values of
Δ, respectively). Microemulsions carry the
requirement of a particular characteristic size, and can only effectively stabilize lipid domains
smaller than that size. If the set cluster of Lo components is larger than the microemulsion
length scale, then there is actually exclusion of other Lo components from the cluster (Figure
6d). In contrast, if the membrane sits near an Ising critical point, the consequent lipid rafts are
24
stabilized at all length scales, never excluding other Lo components. If the membrane indeed
exists as a microemulsion, then in principle it should be possible to experimentally exceed the
correct length scale, and cause a reversal of the lipid mediated signaling. To our knowledge,
this exact experiment has not been carried out, and may remain challenging to implement.
However, in mast cells, a structurally defined ligand with spacing 13 nm has been
studied,52 and the resulting large receptor spacing lowers, but does not eliminate, the
signaling response. This suggests that, if the mast cell signaling response relies on a
microemulsion-mediated kinase recruitment, that microemulsion length scale must be larger
than 13 nm.
One argument sometimes used in favor of microemulsions is that they are easier to
achieve, requiring less cell-directed tuning of the membrane. However, our phase diagram
points to an additional complication: the cell not only has to tune the membrane composition
to a microemulsion, but also must tune the length scale to the characteristic size necessary
for the correct biological function, which may be highly variable, depending on the signaling
pathway and components involved.
What about the actin cytoskeleton? It is widely thought that cortical actin couples to the
membrane, forming "corrals" that add further complexity to the heterogeneity of the
membrane. However, in many ways this does not affect our conclusions, as typical size
estimates for actin corrals53 are above our simulation size of 50x50 nm (Figure 2c). A small
cluster of Lo-preferring components set within one corral sees a particular membrane
composition, regardless of the corrals boundaries at longer length scales. However, actin
involvement motivates two other considerations. First, we should not ignore the phase-
separated region of the phase diagram (blue, Figures 3-5, S2-S3), as the membrane may
have a phase-separating composition, driven below the diffraction limit only by actin-
mediated partitioning.12,54 We see that a phase-separating membrane would yield a kinase
binding energy similar to the minimum in the microemulsion/critical phase. (Figure 6a,d).
Second, we note that, due to cortical actin, the membrane composition encountered by
receptors might not be the global composition of the membrane. This actin meshwork has
been proposed to preferentially sequester either Lo or Ld lipids,12,34 which would deplete
these from a cluster set in the middle of a corral.
The most striking new discovery from our phase diagram and energy calculations is
the power of a membrane at a tricritical point. Our computations show that near the tricritical
point, the potential binding energy due to lipid rafts increases by a factor of 3 compared to
25
any of the other proposed models: critical point, microemulsion, or phase-separated two-
phase coexistence. This increase in energetic favorability could confer significant advantage
for lateral recruitment in the membrane.
Moreover, the interactions near a tricritical point are long-range in nature, which could
also have important implications for signaling in the natural cell environment. Previous
conformal field theory results have shown that, near an Ising critical point, proteins can
interact via long-range critical Casimir forces.44 Because our method of calculating binding
energies agrees with the results from 44, our system should also exhibit such long-range
interactions near the tricritical point, but with a different power law governing their spatial
decay. The large effect at the tricritical point likely comes from the different critical exponents
of this universality class. In particular, the potential that gives rise to critical Casimir forces
scales with the correlation function g(r), which itself scales as g(r) ~ r -d+2-η near a critical point,
where d is the dimension of our system and η is a universal critical exponent.50 Plugging in
the relevant critical exponents, we see that g(r) scales as r -0.15 at the Ising critical point, but as
r -0.25 at the tricritical point.50,55–58 Thus, at the tricritical point, the critical exponent (η=0.15)
allows attraction between Lo-preferring components to remain stronger at a longer distance,
especially compared to direct chemical bonds or electrostatic interactions (which are
expected to be screened over ~1 nm in the cell), and also longer range than the r-2 attractive
forces mediated by membrane curvature.44
To our knowledge, a tricritical point has not previously been considered as a serious
proposition for the physical basis of lipid rafts, and perhaps for a good reason: achieving
proximity to a tricritical point requires tuning of three relevant parameters, whereas proximity
to an Ising critical point requires only two. In the three-dimensional phase space of the
Blume-Capel model, only a single point is a tricritical point. However, we note that in a many-
component cell membrane with many more than 3 degrees of freedom, there would be more
possibilities for tuning to a tricritical composition. The detailed nature of "lipid rafts" is probably
quite variable even within a single functional cell membrane, and localized tuning may be
possible for a particular signaling purpose. Furthermore, we argue that if effective lipid rafts
provide a strong enough evolutionary advantage for the cell to respond appropriately to
environmental stimuli, it might be to the cell's advantage to maintain a tricritical composition
(at least locally), and gain the massive improvement in lipid raft energetics that results.
Conversely, the optimal lipid raft strength for signaling to be appropriately regulated in the cell
might be weaker than what is generated by the tricritical point, in which case we would expect
26
the membrane to exist in one of the other phase states explored in this study.
It is also reasonable to ask whether lipid rafts could facilitate interactions between Lo-
preferring components that lead to the formation of the cluster itself. This was not the case
we considered for mast cells, in which clustering was due to physical cross-linking of the IgE-
FcεRI by antigen. In T cell receptor signaling, for example, clusters form in the absence of
cross-linking by a mechanism that remains unclear.59 The Ising critical point or microemulsion
binding energy of -0.6 kBT would not be sufficient to cause clustering of individual receptors;
this requires considerably stronger interactions. We previously performed calculations and
simulations based on the formulas for Casimir forces given in 44 and concluded that these
forces, at an Ising critical point, are not large enough to mediate receptor clustering (Milka
Doktorova and Eshan Mitra, unpublished observations). However, we now note that the
stronger binding energies found near the tricritical point may be sufficient to mediate receptor
co-clustering, independently of external agent.
We further note that the concept of a membrane at a tricritical point is not inconsistent
with observations of GPMVs showing ordinary Ising critical exponents.21 We argue that a
membrane might exhibit tricritical behavior at short length scales and Ising critical behavior at
the longer length scale accessible with current experimental techniques. This hypothesis can
be formalized using renormalization group (RG) theory, a tool for describing how the
observed behavior of a system changes due to coarse-graining. Here, coarse-graining
corresponds to the loss in resolution when a membrane is observed with a diffraction-limited
microscope. Certain points in parameter space are RG fixed points, which are unaffected by
coarse-graining (i.e. look the same at different length scales). Other points, under RG coarse-
graining "flow" towards or away from the fixed points (Figure S10). The 2D tricritical fixed
point and Ising critical fixed point are two examples of such RG fixed points, with systems
tending to flow from tricritical to Ising behavior under coarse-graining. As seen in Figure S10,
physical systems that flow near to the tricritical point will show tricritical behavior on length
scales relevant for protein organization, but could then flow away to Ising behavior on the
longer length scales observed in GPMV studies.f
While our work with this lattice model has been useful in addressing many hypotheses
on lipid organization (and proposing a new, tricritical possibility), it has some limitations. In
f
Indeed, the phase diagram of a physical system near a critical point echoes the flow diagram near the
corresponding renormalization-group fixed point (the irrelevant contracting directions only making analytic
changes in the phase boundaries), leading to a common conflation of the two (adding 'flow' arrows to the
boundaries in experimental phase diagrams).
27
particular, this is a thermodynamic model, operating under the assumption of a steady state.
Kinetic hypotheses about lipid organization, such as active actin remodeling,60 would require
a different theoretical framework in order to compare to the cases that we have explored.
However, our neural network-based methods should allow similar morphological
classification. Moreover, while it is possible for active processes to be described by Ising
critical behavior,61 studies on GPMVs isolated from cells21 show that these membranes
remain close to an Ising critical point even after any active processes have likely been
disrupted in sample preparation.
Another future direction for this theoretical approach is to convert the phase diagrams
using external fields H and Δ into diagrams based on the concentration of each component.
We chose to use a model with fixed external fields and variable composition to enable
efficient simulations on small system sizes, and to easily compare with existing theory
literature. These external fields could be converted to the corresponding compositions of
each component, transforming the phase diagram to one of fixed compositions. This would
allow more direct comparison to experimental phase diagrams of model membranes such as
in 19.
Acknowledgements
EM, DH, and BB were supported by National Institutes of Health grants R01-AI018306
followed by R01-GM117552, and their contributions further benefited from participation in the
HHMI/MBL Summer Institute supported by an HCIA award. SW was supported by the
Department of Defense through the National Defense Science Engineering Graduate
Fellowship (NDSEG) Program. JPS was supported by the National Science Foundation
grants DMR-1312160 and DMR-1719490. EM was additionally supported by the National
Institutes of Health under the Ruth L. Kirschstein National Research Service Award
(2T32GM008267) from the National Institute of General Medical Sciences.
We are grateful to Frank Zhang for discussions on the use of neural networks for
phase diagram prediction, to Archishman Raju, Colin Clement, and Benjamin Machta for
discussions on critical phenomena and scaling analysis, and to Gerald Feigenson for
discussions on lipid membranes.
28
References
(1) Schmid, F. Physical Mechanisms of Micro- and Nanodomain Formation in Multicomponent
Lipid Membranes. Biochim. Biophys. Acta - Biomembr. 2017, 1859 (4), 509–528.
(2) Léonard, C.; Alsteens, D.; Dumitru, A. C.; Mingeot-Leclercq, M.-P.; Tyteca, D. Lipid
Domains and Membrane (Re)Shaping: From Biophysics to Biology. In The Biophysics
of Cell Membranes: Biological Consequences; Epand, R. M., Ruysschaert, J.-M., Eds.;
Springer Singapore: Singapore, 2017; pp 121–175.
(3) Munro, S. Lipid Rafts. Cell 2003, 115 (4), 377–388.
(4) Pike, L. J. Rafts Defined: A Report on the Keystone Symposium on Lipid Rafts and Cell
Function. J. Lipid Res. 2006, 47 (7), 1597–1598.
(5) Sengupta, P.; Holowka, D.; Baird, B. Fluorescence Resonance Energy Transfer between
Lipid Probes Detects Nanoscopic Heterogeneity in the Plasma Membrane of Live Cells.
Biophys. J. 2007, 92 (10), 3564–3574.
(6) Lingwood, D.; Simons, K. Lipid Rafts as a Membrane-Organizing Principle. Science 2010,
327 (5961), 46–50.
(7) Schick, M. Theories of Equilibrium Inhomogeneous Fluids
http://faculty.washington.edu/schick/Abstracts/sens-book.pdf (accessed Jun 6, 2017).
(8) Dietrich, C.; Bagatolli, L. a.; Volovyk, Z. N.; Thompson, N. L.; Levi, M.; Jacobson, K.;
Gratton, E. Lipid Rafts Reconstituted in Model Membranes. Biophys. J. 2001, 80 (3),
1417–1428.
(9) Veatch, S. L.; Keller, S. L. Seeing Spots: Complex Phase Behavior in Simple Membranes.
Biochim. Biophys. Acta - Mol. Cell Res. 2005, 1746 (3), 172–185.
(10) Veatch, S. L.; Keller, S. L. Separation of Liquid Phases in Giant Vesicles of Ternary
Mixtures of Phospholipids and Cholesterol. Biophys. J. 2003, 85 (5), 3074–3083.
(11) Baumgart, T.; Hammond, A. T.; Sengupta, P.; Hess, S. T.; Holowka, D. A.; Baird, B. A.;
Webb, W. W. Large-Scale Fluid/fluid Phase Separation of Proteins and Lipids in Giant
Plasma Membrane Vesicles. PNAS 2007, 104 (9), 3165–3170.
(12) Machta, B. B.; Papanikolaou, S.; Sethna, J. P.; Veatch, S. L. Minimal Model of Plasma
Membrane Heterogeneity Requires Coupling Cortical Actin to Criticality. Biophys. J.
2011, 100 (7), 1668–1677.
(13) Swamy, M. J.; Ciani, L.; Ge, M.; Smith, A. K.; Holowka, D.; Baird, B.; Freed, J. H.
Coexisting Domains in the Plasma Membranes of Live Cells Characterized by Spin-
Label ESR Spectroscopy. Biophys. J. 2006, 90 (12), 4452–4465.
(14) Simons, K.; Gerl, M. J. Revitalizing Membrane Rafts: New Tools and Insights. Nat. Rev.
Mol. Cell Biol. 2010, 11 (10), 688–699.
29
(15) Holowka, D.; Gosse, J. A.; Hammond, A. T.; Han, X.; Sengupta, P.; Smith, N. L.;
Wagenknecht-Wiesner, A.; Wu, M.; Young, R. M.; Baird, B. Lipid Segregation and IgE
Receptor Signaling: A Decade of Progress. Biochim. Biophys. Acta 2005, 1746 (3),
252–259.
(16) Sezgin, E.; Levental, I.; Mayor, S.; Eggeling, C. The Mystery of Membrane Organization:
Composition, Regulation and Roles of Lipid Rafts. Nat. Rev. Mol. Cell Biol. 2017, 18
(6), 361–374.
(17) Chini, B.; Parenti, M. G-Protein Coupled Receptors in Lipid Rafts and Caveolae: How,
When and Why Do They Go There? J. Mol. Endocrinol. 2004, 32 (2), 325–338.
(18) Nussinov, R.; Jang, H.; Tsai, C. J. Oligomerization and Nanocluster Organization Render
Specificity. Biol. Rev. 2015, 90 (2), 587–598.
(19) Feigenson, G. W. Phase Diagrams and Lipid Domains in Multicomponent Lipid Bilayer
Mixtures. Biochim. Biophys. Acta 2009, 1788 (1), 47–52.
(20) Konyakhina, T. M.; Feigenson, G. W. Phase Diagram of a Polyunsaturated Lipid Mixture:
Brain sphingomyelin/1-Stearoyl-2-Docosahexaenoyl-Sn-Glycero-3-
Phosphocholine/cholesterol. Biochim. Biophys. Acta - Biomembr. 2016, 1858 (1), 153–
161.
(21) Veatch, S. L.; Cicuta, P.; Sengupta, P.; Honerkamp-Smith, A.; Holowka, D.; Baird, B.
Critical Fluctuations in Plasma Membrane Vesicles. ACS Chem. Biol. 2008, 3 (5), 287–
293.
(22) Stanich, C. A.; Honerkamp-Smith, A. R.; Putzel, G. G.; Warth, C. S.; Lamprecht, A. K.;
Mandal, P.; Mann, E.; Hua, T.-A. D.; Keller, S. L. Coarsening Dynamics of Domains in
Lipid Membranes. Biophys. J. 2013, 105 (2), 444–454.
(23) Konyakhina, T. M.; Goh, S. L.; Amazon, J.; Heberle, F. A.; Wu, J.; Feigenson, G. W.
Control of a Nanoscopic-to-Macroscopic Transition: Modulated Phases in Four-
Component DSPC/DOPC/POPC/Chol Giant Unilamellar Vesicles. Biophys. J. 2011,
101 (2), L8–L10.
(24) Levental, I.; Veatch, S. L. The Continuing Mystery of Lipid Rafts. J. Mol. Biol. 2016, 428
(24), 4749–4764.
(25) Schick, M. Membrane Heterogeneity: Manifestation of a Curvature-Induced
Microemulsion. Phys. Rev. E 2012, 85 (3), 031902.
(26) Sadeghi, S.; Müller, M.; Vink, R. L. C. Raft Formation in Lipid Bilayers Coupled to
Curvature. Biophys. J. 2014, 107 (7), 1591–1600.
(27) Amazon, J. J.; Goh, S. L.; Feigenson, G. W. Competition between Line Tension and
Curvature Stabilizes Modulated Phase Patterns on the Surface of Giant Unilamellar
30
Vesicles: A Simulation Study. Phys. Rev. E 2013, 87 (2), 22708.
(28) Palmieri, B.; Safran, S. A. Hybrid Lipids Increase the Probability of Fluctuating
Nanodomains in Mixed Membranes. Langmuir 2013, 29 (17), 5246–5261.
(29) Honerkamp-Smith, A. R.; Veatch, S. L.; Keller, S. L. An Introduction to Critical Points for
Biophysicists; Observations of Compositional Heterogeneity in Lipid Membranes.
Biochim. Biophys. Acta 2009, 1788 (1), 53–63.
(30) Blank, U.; Rivera, J. The Ins and Outs of IgE-Dependent Mast-Cell Exocytosis. Trends
Immunol. 2004, 25 (5), 266–273.
(31) Gilfillan, A. M.; Rivera, J. The Tyrosine Kinase Network Regulating Mast Cell Activation.
Immunol. Rev. 2009, 228 (1), 149–169.
(32) Rivera, J.; Gilfillan, A. M. Molecular Regulation of Mast Cell Activation. J. Allergy Clin.
Immunol. 2006, 117 (6), 1214–1225.
(33) Holowka, D.; Baird, B. Roles for Lipid Heterogeneity in Immunoreceptor Signaling.
Biochim. Biophys. Acta - Mol. Cell Biol. Lipids 2016, 1861 (8, Part B), 830–836.
(34) Shelby, S. A.; Veatch, S. L.; Holowka, D. A.; Baird, B. A. Functional Nanoscale Coupling
of Lyn Kinase with IgE-FcεRI Is Restricted by the Actin Cytoskeleton in Early Antigen-
Stimulated Signaling. Mol. Biol. Cell 2016, 27 (22), 3645–3658.
(35) Gompper, G.; Schick, M. Lattice Model of Microemulsions: The Effect of Fluctuations in
One and Two Dimensions. Phys. Rev. A 1990, 42 (4), 2137–2149.
(36) Gompper, G.; Schick, M. Lattice Model of Microemulsions. Phys. Rev. B 1990, 41 (13),
9148–9162.
(37) Blume, M.; Emery, V. J.; Griffiths, R. B. Ising Model for the Λ Transition and Phase
Separation in He3-He4 Mixtures. Phys. Rev. A 1971, 4 (3), 1071–1077.
(38) Blume, M. Theory of the First-Order Magnetic Phase Change in UO2. Phys. Rev. 1966,
141 (2), 517–524.
(39) Capel, H. W. On the Possibility of First-Order Transitions in Ising Systems of Triplet Ions
with Zero-Field Splitting. Physica 1966, 32 (5), 966–988.
(40) Krizhevsky, A.; Sutskever, I.; Hinton, G. E. ImageNet Classification with Deep
Convolutional Neural Networks; Pereira, F., Burges, C. J. C., Bottou, L., Weinberger, K.
Q., Eds.; Curran Associates, Inc., 2012.
(41) Carrasquilla, J.; Melko, R. G. Machine Learning Phases of Matter. Nat. Phys. 2017, 13
(5), 431–434.
(42) Wang, L. Discovering Phase Transitions with Unsupervised Learning. Phys. Rev. B -
Condens. Matter Mater. Phys. 2016, 94 (19), 195105.
31
(43) Gompper, G.; Schick, M. Lattice Model of Microemultions. Phys. Rev. B 1990, 41 (13),
9148–9162.
(44) Machta, B. B.; Veatch, S. L.; Sethna, J. P. Critical Casimir Forces in Cellular Membranes.
Phys. Rev. Lett. 2012, 109 (13), 1–5.
(45) Nielsen, M. A. Neural Networks and Deep Learning neuralnetworksanddeeplearning.com
(accessed Jan 1, 2017).
(46) Bennett, C. H. Efficient Estimation of Free Energy Differences from Monte Carlo Data. J.
Comput. Phys. 1976, 22, 245–268.
(47) Metropolis, N.; Rosenbluth, A. W.; Rosenbluth, M. N.; Teller, A. H.; Teller, E. Equation of
State Calculations by Fast Computing Machines. J. Chem. Phys. 1953, 21 (6), 1087–
1092.
(48) Jarzynski, C. Nonequilibrium Equality for Free Energy Differences. Phys. Rev. Lett.
1997, 78 (14), 2690–2693.
(49) Beale, P. D. Finite-Size Scaling Study of the Two-Dimensional Blume-Capel Model.
Phys. Rev. B 1986, 33 (3), 1717–1720.
(50) Cardy, J. Scaling and Renormalization in Statistical Physics; Cambridge University
Press: New York, New York, USA, 1996.
(51) Palmieri, B.; Grant, M.; Safran, S. A. Prediction of the Dependence of the Line Tension
on the Composition of Linactants and the Temperature in Phase Separated
Membranes. Langmuir 2014, 30 (39), 11734–11745.
(52) Sil, D.; Lee, J. B.; Luo, D.; Holowka, D.; Baird, B. Trivalent Ligands with Rigid DNA
Spacers Reveal Structural Requirements for IgE Receptor Signaling in RBL Mast Cells.
ACS Chem. Biol. 2007, 2 (10), 674–684.
(53) Kusumi, A.; Fujiwara, T. K.; Morone, N.; Yoshida, K. J.; Chadda, R.; Xie, M.; Kasai, R. S.;
Suzuki, K. G. N. Membrane Mechanisms for Signal Transduction: The Coupling of the
Meso-Scale Raft Domains to Membrane-Skeleton-Induced Compartments and
Dynamic Protein Complexes. Semin. Cell Dev. Biol. 2012, 23 (2), 126–144.
(54) Honigmann, A.; Sadeghi, S.; Keller, J.; Hell, S. W.; Eggeling, C.; Vink, R. A Lipid Bound
Actin Meshwork Organizes Liquid Phase Separation in Model Membranes. eLife 2014,
3, e01671.
(55) Nienhuis, B.; Berker, A. N.; Riedel, E. K.; Schick, M. First- and Second-Order Phase
Transitions in Potts Models: Renormalization-Group Solution. Phys. Rev. Lett. 1979, 43
(11), 737–740.
(56) Pearson, R. B. Conjecture for the Extended Potts Model Magnetic Eigenvalue. Phys.
Rev. B 1980, 22 (5), 2579–2580.
32
(57) Nienhuis, B.; Warnaar, S. O.; Blote, H. W. J. Exact Multicritical Behaviour of the Potts
Model. J. Phys. A. Math. Gen. 1993, 26 (3), 477.
(58) Kwak, W.; Jeong, J.; Lee, J.; Kim, D.-H. First-Order Phase Transition and Tricritical
Scaling Behavior of the Blume-Capel Model: A Wang-Landau Sampling Approach.
Phys. Rev. E. Stat. Nonlin. Soft Matter Phys. 2015, 92 (2-1), 22134.
(59) Sherman, E.; Barr, V.; Samelson, L. E. Super-Resolution Characterization of TCR-
Dependent Signaling Clusters. Immunol. Rev. 2013, 251 (1), 21–35.
(60) Rao, M.; Mayor, S. Active Organization of Membrane Constituents in Living Cells. Curr.
Opin. Cell Biol. 2014, 29, 126–132.
(61) Noble, A. E.; Machta, J.; Hastings, A. Emergent Long-Range Synchronization of
Oscillating Ecological Populations without External Forcing Described by Ising
Universality. Nat. Commun. 2015, 6, 6664.
33
Supporting Information
Figure S1: Training the neural network. (a) Each circle at specified (Δ/J, T/J) coordinates
represents a parameter set at which training data were acquired. The color indicates the
human-generated classification of the phase at that point. These points are overlaid on the
final phase diagram computed by the neural network. (b-c) Phase diagrams generated by two
types of neural network. (b) is trained on averages of correlated Monte Carlo snapshots to
identify all six phases, while (c) is trained on individual Monte Carlo snapshots to identify four
S1
phases. Our final phase diagram is based mostly on (b), but we take the classification of the
gray phase from (c) due to its high confidence in that phase. (d) Example simulation images
taken near the Fluid / Gray boundary (Fluid at T/J = 1.49 , Δ/J = 6.2 ; Gray at T/J = 1.49 , Δ/J
= 6.8), demonstrating the one case in which working with snapshots (phase diagram (c)) is
advantageous compared to working with averages (phase diagram (d)). Note that the
snapshots shown are easily distinguishable as two different phases, but the averages shown
look more similar to each other.
S2
Figure S2: Cross-sections of the phase diagram in which H/J=0, K/J=0, and L/J = 1 or 2.
Colors have the same meaning as in Figure 3 and Figure S1.
S3
Figure S3: Additional cross-sections of the phase diagram at positive H. Colors have the
same meaning as in Figure 3 and Figure S1.
S4
Figure S4: Free energy change (binding energy) due to moving two white pixels to a spacing
of 3 lattice units apart. Simulations were run on a system similar to that shown in Figure 5, but
with the box on the left containing, instead of the set cluster of 3 IgE receptors (each
represented by 12 pixels), a single fixed white pixel. Each diamond indicates the free energy
calculated by Bennett's method at that point in the phase diagram. (a) Binding energy in the
Blume-Capel model (H=0, K=0, L=0) cross-section. (b) Inset of (a) in the area of the tricritical
point (black box in (a)). Similar to the more complicated case of clustered receptors (Figure
6a-b), the binding energy is considerably higher close to the tricritical point. (c) Binding energy
S5
in a cross-section that includes microemulsions (K/J = 2, L/J = 3). Similar to the more
complicated case (Figure 6d), certain microemulsions give a positive binding energy.
We provide data for this case, which is simpler than that shown in the main text, in the hopes
that it will prove useful for future theoretical work related to this model. In particular, we note
that it may be possible develop a universal scaling theory to describe the increase in binding
energy magnitude as the tricritical point is approached.
S6
Figure S5: Alternate method for calculating kinase binding energy. We use the method
described in 1. (a-b) Example simulation at the tricritical point, in which the kinase sits at a
position 95 nm from the cluster. As in Figure 5, receptors (red) and kinase (green) were held
fixed as white, and the rest of the lattice was Monte Carlo sampled. (a) shows one simulation
snapshot, and (b) shows the average of all snapshots acquired. The simulation was repeated
S7
at every possible kinase position from 0 to 95, stepwise, in order to generate (c), the profile
showing the free energy of the kinase in every position. The position axis in (c) corresponds
to the axis shown in (a). The free energy at position 3 nm of ~ -1.4 kBT is in good agreement
with the computationally cheaper method used in the rest of the study (Figure 6, Figure S4).
S8
Figure S6: Detailed analysis of three selected points in the H/J=0, K/J=2, L/J=3 cross-
section. (a) Correlation functions between white pixels at the selected points. The 2D
correlation functions G are calculated according to the formula
G= 1
ρ2
FFT −1(FFT ( A)2)
FFT −1(FFT (M )2)
where A is a 2D array with value 1 if the σ value at the corresponding pixel of the snapshot is
1 (white pixel), or 0 otherwise, M is an array of all 1's of the same size as A, ρ is the density of
S9
white pixels (calculated as the sum of A divided by the sum of M), and FFT represents the
fast Fourier transform. A correlation function value greater than 1 indicates positive
correlation, 1 indicates no correlation, and less than 1 indicates anticorrelation. The function
plotted here is the average transect of G along the horizontal (1,0) and vertical (0,1) directions
(used instead of a radial average because G is not always radially symmetric). Correlation
functions are averaged over 2000 snapshots of size 200x200 per point.
One criterion to define a microemulsion is that this correlation function must have a local
minimum, rather than decrease monotonically. By this criterion, T=2.01, Δ=2.0 is not a
microemulsion, as the correlation function decreases monotonically. T=1.35, Δ=4.85 is a
microemulsion with length scale ~ 10 because the first local minimum occurs at lattice
distance 10, and T=1.25, Δ=5.7 is a microemulsion with length scale ~ 4 because the
(considerably stronger) first local minimum occurs at lattice distance 4. (b) The analysis using
correlation functions gives results consistent with a visual inspection of snapshots at the same
three points, which appear as critical fluctuations (left), a long length-scale microemulsion
(center), and a short length-scale microemulsion (right).
S10
Figure S7: Neural network computation of the Ising model phase diagram. Graphs show the
outputs of the neural network corresponding to each phase, indicating the network's level of
confidence that the system represents that phase. At each temperature value, the network's
prediction is the phase for which the network output is the highest. The true Ising transition
temperature is indicated by the dashed line; an accurate network should give "Separated" as
the highest output below this temperature, and "Critical" or "Fluid" as the highest output above
this temperature. (a) Output of the neural network used in the rest of this study. The predicted
transition temperature is close to, but slightly above, the true transition temperature. (b)
Output of a neural network specialized for the Ising model only. This network was trained on
examples from the Ising model, and only distinguishes between the fluid and phase-
separated states. This network's predicted transition temperature almost exactly matches the
true transition temperature.
S11
Figure S8: Comparison to other methods for calculation of the Blume-Capel phase diagram.
The neural network's phase diagram follows the same key as in Figure 2. Overlaid in gray is
the phase boundary as determined by finite size scaling in 2. The dashed line represents a
critical phase transition, solid represents a first-order transition, and the transition from
dashed to solid is the tricritical point. Dotted line in blue is the critical line as determined by
mean field theory in 3.
S12
Figure S9: Comparing phase diagram results from our study to the original study. Gray
overlays are reproduced from the original study.4 Solid lines indicate first order phase
transitions, dashed lines indicate critical phase transitions, and dotted lines are Lifshitz lines.
S13
Figure S10: Schematic of renormalization group flows and crossover behavior. The operation
of coarse-graining leads to flows through the space of possible models (i.e. a space whose
different directions correspond to different model parameters), a process formalized by the
renormalization group (RG). Fixed points in this model space are models that look the same
at all length scales, and correspond to different classes of critical phenomena. (a) gives an
example of RG flows in the familiar Ising model. SIsing is the Ising critical fixed point; under
coarse-graining, models flow towards it along black lines and away along gray lines. For the
Ising model, temperature (vertical, dotted red arrow) is crucial in determining model behavior.
Above the critical temperature, Tc, as we coarse-grain our model (RG Flow 1), it flows to a
mixed (uniform) fluid state. Below Tc (RG Flow 2), the model flows toward a phase-separated
state. Importantly, when the models in either RG Flow 1 or 2 pass near SIsing, they exhibit
Ising-like properties, and coarse-graining to length scales of 20 nm is typically sufficient to
produce critical behavior. (b) In the presence of more than one critical fixed point, a model's
trajectory through RG flows can pass near-and thus be influenced by-the multiple different
fixed points. In this schematic, we imagine the trajectory of a membrane model in such a
space (gray dashed arrow), where STri and SIsing are the 2D tricritical and 2D Ising critical fixed
S14
points (respectively). At T=Tc, where Tc is the Ising critical temperature, the model flows
directly to the Ising fixed point. The transition from tricritical-like to Ising-like behavior is
referred to as crossover. (c) In the same space as (b), above Tc, our model can flow along a
complicated trajectory (red dashed arrow line). Note that here we have included a 3rd
dimension-temperature, which acts as in (a). At short length scales (~20 nm), our model is
near the tricritical fixed point, and will thus obey tricritical behavior. As we coarse-grain further,
our model flows away toward the Ising critical point, so that the behavior of the model at
intermediate length scales (~200 nm) is Ising-like. As we coarse-grain even further, the model
flows away to some non-critical fixed point.
S15
Supporting References
(1) Machta, B. B.; Veatch, S. L.; Sethna, J. P. Critical Casimir Forces in Cellular
Membranes. Phys. Rev. Lett. 2012, 109 (13), 138101.
(2) Beale, P. D. Finite-Size Scaling Study of the Two-Dimensional Blume-Capel Model.
Phys. Rev. B 1986, 33 (3), 1717–1720.
(3) Blume, M.; Emery, V. J.; Griffiths, R. B. Ising Model for the λ Transition and Phase
Separation in He3-He4 Mixtures. Phys. Rev. A 1971, 4 (3), 1071–1077.
(4) Gompper, G.; Schick, M. Lattice Model of Microemulsions: The Effect of Fluctuations in
One and Two Dimensions. Phys. Rev. A 1990, 42 (4), 2137–2149.
S16
|
1805.12516 | 5 | 1805 | 2019-02-13T16:34:50 | Available energy fluxes drive a transition in the diversity, stability, and functional structure of microbial communities | [
"physics.bio-ph",
"cond-mat.stat-mech",
"q-bio.PE"
] | A fundamental goal of microbial ecology is to understand what determines the diversity, stability, and structure of microbial ecosystems. The microbial context poses special conceptual challenges because of the strong mutual influences between the microbes and their chemical environment through the consumption and production of metabolites. By analyzing a generalized consumer resource model that explicitly includes cross-feeding, stochastic colonization, and thermodynamics, we show that complex microbial communities generically exhibit a transition as a function of available energy fluxes from a "resource-limited" regime where community structure and stability is shaped by energetic and metabolic considerations to a diverse regime where the dominant force shaping microbial communities is the overlap between species' consumption preferences. These two regimes have distinct species abundance patterns, different functional profiles, and respond differently to environmental perturbations. Our model reproduces large-scale ecological patterns observed across multiple experimental settings such as nestedness and differential beta diversity patterns along energy gradients. We discuss the experimental implications of our results and possible connections with disorder-induced phase transitions in statistical physics. | physics.bio-ph | physics | Available energy fluxes drive a transition in the diversity, stability, and functional
structure of microbial communities
Robert Marsland III,∗ Wenping Cui,† Joshua Goldford,‡ Alvaro Sanchez,§ Kirill Korolev, and Pankaj Mehta
Department of Physics, Boston University, Boston, MA
(Dated: February 14, 2019)
A fundamental goal of microbial ecology is to understand what determines the diversity, sta-
bility, and structure of microbial ecosystems. The microbial context poses special conceptual chal-
lenges because of the strong mutual influences between the microbes and their chemical environment
through the consumption and production of metabolites. By analyzing a generalized consumer re-
source model that explicitly includes cross-feeding, stochastic colonization, and thermodynamics, we
show that complex microbial communities generically exhibit a transition as a function of available
energy fluxes from a "resource-limited" regime where community structure and stability is shaped
by energetic and metabolic considerations to a diverse regime where the dominant force shaping
microbial communities is the overlap between species' consumption preferences. These two regimes
have distinct species abundance patterns, different functional profiles, and respond differently to
environmental perturbations. Our model reproduces large-scale ecological patterns observed across
multiple experimental settings such as nestedness and differential beta diversity patterns along en-
ergy gradients. We discuss the experimental implications of our results and possible connections
with disorder-induced phase transitions in statistical physics.
INTRODUCTION
Microbial communities inhabit every corner of our
planet, from our own nutrient-rich guts to the remote
depths of the ocean floor. Different environments har-
bor very different levels of microbial diversity:
in some
samples of non-saline water at mild temperature and pH,
nearly 3,000 coexisting types of bacteria can be detected,
whereas at ambient temperatures warmer than 40◦ C,
most cataloged samples contain fewer than 100 distinct
variants [49]. The functional structure of these commu-
nities is also highly variable, with functional traits often
reflecting the environment in which the communities are
found [27, 49]. A central goal of microbial community
ecology is to understand how these variations in diver-
sity, stability and functional structure [55] arise from an
interplay of environmental factors such as energy and re-
source availability [15, 34] and ecological processes such
as competition [10, 21, 32, 35] and stochastic colonization
[9, 28, 30, 53].
This endeavor is complicated by the fact that microbes
dramatically modify their abiotic environments through
consumption and secretion of organic and inorganic com-
pounds. This happens on a global scale, as in the Great
Oxidation Event two billion years ago [5, 45], and also
on smaller scales relevant to agriculture, industry and
medicine.
In this sense, every microbe is an "ecosys-
tem engineer" [29]. Metabolic modeling and experiments
suggests that metabolically-mediated syntrophic interac-
tions should be a generic feature of microbial ecology
∗ [email protected]
† Department of Physics, Boston College, Chestnut Hill, MA
‡ Bioinformatics Program, Boston University, Boston, MA
§ Department of Ecology and Evolutionary Biology, Yale Univer-
sity, New Haven, CT
[23, 26, 58] and that complex microbial communities can
self-organize even in constant environments with no spa-
tial structure or predation [19, 23]. For these reasons,
there has been significant interest in developing new mod-
els for community assembly suited to the microbial set-
ting [8, 24, 25, 48, 51].
Here, we present a statistical physics-inspired con-
sumer resource model for microbial community assembly
that builds upon the simple model introduced in [23] and
explicitly includes energetic fluxes, stochastic coloniza-
tion, syntrophy, and resource competition. We focus on
modeling complex communities with many species and
metabolites. By necessity, any mathematical model of
such a large, diverse ecosystem will contain thousands of
parameters that are hard to measure. To circumvent this
problem, we take a statistical physics approach where
all consumer preferences and metabolic parameters are
drawn from random distributions.
This approach to modeling complex systems has its
root in the pioneering work of Wigner on the spectrum
of heavy nuclei [56] and was adapted by May to ecological
settings [37]. Recently, there has been a renewed inter-
est in using these ideas to understand complex systems
in both many-body physics (reviewed in [11]) and com-
munity assembly [1, 4, 6, 7, 13, 18, 22, 23, 30, 51]. The
key insight underlying this approach is that generic and
reproducible large-scale patterns observed across multi-
ple settings likely reflect typical properties, rather than
fine tuned features of any particular realization or com-
munity. Consistent with this idea, it was recently shown
that a generalized consumer resource model with random
parameters can reproduce many of the patterns observed
in experiments where natural communities were grown in
synthetic minimal environments [23].
In this paper, we ask how varying the energy flux
into an ecosystem and the amount of cross-feeding af-
fects microbial community assembly. We find that the
9
1
0
2
b
e
F
3
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
5
v
6
1
5
2
1
.
5
0
8
1
:
v
i
X
r
a
2
FIG. 1. Microbial communities engineer complex chemical environments using a single energy source.
(A)
Schematic of microbe-mediated energy fluxes in the Thermodynamic Microbial Consumer Resource Model. Each cell of species
i (= 1, 2, . . . S) supplies itself with energy through import of resources, generating an incoming energy flux J in
iα for each resource
type α (= 1, 2, . . . M ). A fraction lα of this energy leaks back into the environment in the form of metabolic byproducts, with
each byproduct type carrying an outgoing energy flux J out
, is used to replicate
the cell. (B) Each species is defined by a vector of consumer preferences that encode its capacity for harvesting energy from
each resource type. These vectors comprise a consumer matrix ciα. (C) A regional pool of species is randomly generated, and
communities are initialized with random subsets of these species to simulate stochastic colonization. (D) Each community is
supplied with a constant flux κ0 of a single resource type (α = 0), and all resources are continuously diluted at a fixed rate
−1
τ
R . (E) Consumer populations Ni and resource concentrations Rα as a function of time for two realizations of this model,
with low (l = 0.001) and high (l = 0.8) levels of uniform metabolic leakage (see Appendix B for parameters).
iα. The remaining energy, J grow
iβ =(cid:80)
α lαDβαJ in
i
resulting communities generically fall into two distinct
regimes, characterized by qualitative differences in their
community-level metabolic networks, functional struc-
tures, responses to environmental perturbations, and
large-scale biodiversity patterns. We show our model pre-
dictions are consistent with data from the Tara Oceans
database [47] and the Earth Microbiome Project [49], and
propose feasible experimental tests using synthetic com-
munities.
METHODS
The starting point for our analysis is a new model that
adapts MacArthur's Consumer Resource Model [35] to
the microbial context by including energetics, stochas-
tic colonization, and the exchange and consumption of
metabolites. We consider the population dynamics of S
species of consumers (e.g., microbes) competing for M
types of substitutable resources. We are interested in
large, diverse ecosystems where S, M (cid:29) 1. A schematic
summarizing our model is shown in Figure 1.
A natural setting for considering substitutable re-
sources is when all essential biomass components are sup-
plied in excess, and the limiting factor for growth is the
supply of usable energy. In this context, one only needs
to keep track of resources from which energy can be har-
vested. All other nutrients are included implicitly, under
the assumption that some of the energy budget is used to
import whatever materials are required for growth and
reproduction. Terminal waste products from which no
more energy can be extracted are likewise treated implic-
itly, and are not included among the M resource types.
In our model, the rate at which an individual of species
i harvests energy from resource α depends on the resource
concentration Rα as well as on the consumer's vector of
resource preferences ciα through the expression:
J in
iα = wασ(ciαRα),
(1)
where σ(x) encodes the functional response and has units
of mass/time, while wα is the energy density of resource
α with units of energy/mass. In the microbial context the
consumer preferences ciα can be interpreted as expression
levels of transporters for each of the resources.
In the
main text, we focus on Type-I responses where σ(x) = x,
and we set wα = 1 for all α, but most of our results
still hold when σ(x) is a Monod function or the wα are
randomly sampled, as shown in the SI Appendix.
We model leakage and secretion by letting a fraction
lα of this imported energy return to the environment, so
that the power available to the cell for sustaining growth
is equal to
(cid:88)
α
J grow
i
=
(1 − lα)J in
iα.
(2)
This parameterization guarantees that the community
does not spontaneously generate usable energy in viola-
tion of the Second Law of Thermodynamics. We assume
that a fixed quantity mi of power per cell is required
for maintenance of a steady population of species i, and
that the per-capita growth rate is proportional to the re-
maining energy flux, with proportionality constant gi. In
(B)(C)(A)Low LeakageHigh LeakageJgrowiJini↵Jouti Regional Pool{Ni},{R↵}⌧ 1R(D)(E)D ↵Local Communities(l=0.001)(l=0.8)Resource Conc. R𝛼Consumer Pop. Nitypical experimental conditions, cell death is negligible,
and mi is the energy harvest required for the replication
rate to keep up with the dilution rate. Under these as-
sumptions, the time-evolution of the population size Ni
of species i can be modeled using the equation
(3)
dNi
dt
= giNi [J grow
i
i = (cid:80)
− mi] .
α lαJ in
The leaked energy flux J out
iα from each cell
of species i is partitioned among the M possible resource
types via the biochemical pathways operating within the
cell. We assume that all species share a similar core
metabolism, encoded in a matrix Dβα. Each element
of Dβα specifies the fraction of leaked energy from re-
source α that is released in the form of resource β (note
β Dβα = 1). Thus, in our model
the resources that are excreted into the environment are
intimately coupled to the resources a cell is consuming.
The outgoing energy flux contained in metabolite β is
given by
that by definition, (cid:80)
J out
iβ = wβνout
iβ =
DβαlαJ in
iα.
(4)
(cid:88)
α
(cid:88)
j
The dynamics of the resource concentrations depend on
the incoming and outgoing mass fluxes νin
iα = σ(ciαRα)
and νout
iα , which are related to the energy fluxes via the
energy densities wα. In terms of these quantities, we have
dRα
dt
= hα +
Nj(νout
jα − νin
jα),
(5)
with hα encoding the dynamics of externally supplied
resources. In this manuscript, we focus on the case where
the microbial communities are grown in a chemostat with
a single externally supplied resource α = 0 (Figure 1).
In this case, the resource dynamics can be described by
choosing hα = κα − τ−1
R Rα, with all the κα set to zero
except for κ0. These equations for Ni and Rα, along with
the expressions for J in
iα and J out
iα , completely specify the
ecological dynamics of the model.
This model has been implemented in a freely available
open-source Python package "Community Simula-
tor." The package can be downloaded from https:
//github.com/Emergent-Behaviors-in-Biology/
community-simulator.
RESULTS
To assess the typical community structure and resource
pool stability for ecosystems obeying Equations (1)-(5),
we randomly generated an M×M metabolic matrix Dαβ,
and a binary S× M consumer preference matrix ciα with
S = 200 species and M = 100 resources. Realizations of
the randomly drawn metabolic and consumer preference
matrices are shown in Figure 3 (top) and Figure 5 (left),
respectively.
3
We chose ciα so that each species had 10 preferred re-
source types on average, with ciα = 1, while the rest
of the resources were consumed at a baseline level of
ciα = 0.01. The metabolic matrix Dαβ was sampled
from a Dirichlet distribution, which guarantees that all
the columns sum to 1 as required by the definition of this
parameter. In Appendix C, we show that the qualitative
patterns we observe are unchanged if ciα is drawn from a
Gaussian or Gamma distribution, or if the Dαβ matrix is
made less sparse. The full sampling procedure is detailed
in Appendix A.
We chose our units of energy flux such that the mean
maintenance cost mi over all species in the regional pool
is equal to 1. To break ties between species with simi-
lar consumption profiles, we added a Gaussian random
offset to the mi of each species with standard deviation
0.1. In Figure 19 of the Appendices, we show that these
intrinsic fitness differences do not dominate the ecologi-
cal dynamics, and that many species with relatively high
maintenance costs are able to reach large population sizes
in the steady-state communities.
Finally, we set all the wα equal to 1, and made all the
leakage fractions identical, with lα = l for all α.
To assess the amount of variability in the results, we
initialized 10 different communities by seeding each one
with a random subset of 100 species from the full 200-
species pool. This simulates the stochastic colonization
frequently observed in microbial ecosystems, where the
community composition can randomly vary depending
on the set of microbes this particular local environment
happened to be exposed to [41]. Figure 1 shows typical
dynamical trajectories in the presence of high (l = 0.8)
and extremely low leakage (l = 0.001).
Available energy fluxes drive a transition between a
"resource-limited" and "diverse" regime
Our numerical simulations display a transition between
two qualitatively different community structures as we
vary the externally supplied energy flux w0κ0 and the
leakage/syntrophy l.
In the "thermodynamic limit" of
M, S → ∞, the communities exhibit signatures of a
phase transition analogous to those found in disordered
systems in physics (see Discussion and Appendix D). Fig-
ure 2 shows the effect of this transition on community
diversity at our chosen finite values of S and M . At low
levels of energy flux or syntrophy, the diversity is severely
limited by resource availability. In the limit of high sup-
plied energy flux and high leakage, a maximally diverse
regime is obtained, where the number of surviving species
is limited only by the similarity between consumption
profiles within the regional species pool, in accordance
with classical niche-packing theory [35] as we will discuss
below.
4
The resource-limited regime produces a unidirectional
flow of energy, which is converted from the externally
supplied resource type into an orderly succession of se-
creted resources. For the sparse metabolic matrix shown
in the top row of Figure 3, most resource types also have
extremely small incoming flux vectors in this regime, with
magnitudes less than 1% the size of the largest flux in the
network. The diverse regime displays a qualitatively dif-
ferent structure, where all resources have significant in-
coming fluxes (regardless of the choice of Dαβ), and the
large number of loops in the network makes it impossible
to put the resource types into any definite order. In Ap-
pendix Figure 17, we plot the fraction of samples from
Figure 2 whose (pruned) flux networks are free of cycles,
and confirm that this observation is generic. The dra-
matic contrast between the community-level metabolism
of the two regimes affects many other global features of
the ecosystem, which we will explore in the following sec-
tions.
The two regimes have distinct functional structures
To better understand the behavior of consumers in the
two regimes, we examined the functional traits of mem-
bers of typical communities in each one. In the resource-
limited regime, many surviving species derive most of
their energy directly from the externally supplied re-
source (Figure 4A). In the diverse regime, by contrast,
only a minority of the steady-state community mem-
bers can consume this resource at all, and even these
species receive most of their energy from a diverse ar-
ray of metabolic byproducts (Figure 4B). We quantified
this observation using the Simpson Diversity M eff
of the
incoming resource flux vectors J in
iα, which measures the
effective number of resources consumed by each species,
and is closely related to the inverse participation ratio in
statistical physics. The Simpson Diversity is defined by
i
(cid:19)2(cid:35)−1
(cid:34)(cid:88)
α
(cid:18) J in
iα
J in
i
M eff
i =
,
(7)
i =(cid:80)
i
α J in
where J in
iα is the total incoming energy flux for
each cell of this species. M eff
approaches 1 for species
that obtain the bulk of their energy from a single re-
source type and approaches M when all resource types
are consumed equally.
In the resource-limited regime,
the distribution of these values is sharply peaked around
2. In the diverse regime, the peak is located around 10,
which is the average number of resources with high trans-
porter expression in our binary sampling scheme for ciα.
This shows that most community members in the diverse
regime utilize multiple energy sources, with the incom-
ing flux spread evenly over all resource types they are
capable of consuming.
FIG. 2.
Steady-state richness as a function of
metabolic leakage l and externally supplied energy
flux w0κ0. We generated 200 species, initialized 10 com-
munities of 100 species each from this pool, and ran the dy-
namics to steady state under different combinations of w0κ0
and l (see main text and Appendix B for parameters). (A)
Heat map summarizing all simulations, colored by the av-
erage number of surviving species per steady-state commu-
nity ("Richness"). Slices through the heat map are plotted
in Figure 10 following the Appendices. (B) Community com-
positions are displayed as rank-abundance curves for three
illustrative w0κ0, l combinations (colored by community rich-
ness): (1) "syntrophy-limited" (w0κ0 = 1000, l = 0.1), (2)
"energy-limited" (w0κ0 = 28, l = 0.6) and (3) "similarity-
limited" (w0κ0 = 1000, l = 0.9). The lines are assigned differ-
ent shades for clarity. The first two examples are parts of the
same resource-limited regime, manifesting similar statistical
properties. The plots are truncated at a relative abundance
of 0.5%; see Figure 11 following the Appendices for full data.
The resource-limited and diverse regimes produce
different patterns of energy flux
The difference between the two regimes is most appar-
ent from the perspective of the energy flux networks. Be-
cause our model explicitly accounts for the flow of energy
from one resource type into another, we can compute all
the steady-state fluxes and represent them graphically,
as shown in Figure 3 for some representative examples.
Each node in this network is a resource type, and each
directed edge represents the steady-state flux Jβα of en-
ergy conversion from resource α to resource β, mediated
by one or more syntrophic consumers:
(cid:88)
(cid:88)
Jβα =
NiJ out
iβα = Dβαlα
NiwαciαRα.
(6)
i
i
Resource-Limited RegimeDiverse Regime(B)123(A)5
FIG. 3. Energy flux networks differ in the two regimes. Community-scale energy flux networks are plotted for a
characteristic example from the diverse and resource limited regimes and two different choices of metabolic matrix Dαβ. The
color of each pixel in the heat maps indicates the logarithm (base 10) of the corresponding matrix entry. In the networks,
each node represents one of the M = 100 resource types. Edges represent steady-state energy flux from one resource type
into another, mediated by consumer metabolism and leakage/secretion. The thickness of each edge is proportional to the flux
magnitude, and edges with magnitudes less than 1% of the maximum flux are not displayed. The single node at the top of
each graph is the externally supplied resource, and the rows of nodes at the bottom are resources that are not connected to the
external supply by any flux above the 1% threshold. A topological analysis of the flux networks of all the simulated communities
can be found in Figure 17 of the Appendices.
Responses to resource perturbations differ in the
two regimes
Another important property of microbial ecosystems is
how they respond to environmental perturbations. Pre-
vious theoretical studies have shown that sufficiently di-
verse communities can "pin" the resource concentrations
in their local environment to fixed values, which are in-
dependent of the magnitude of externally supplied fluxes
[44, 48, 50].
In these studies, resource pinning occurs
only when the community saturates the diversity bound
imposed by the principal of competitive exclusion, i.e.
when the number of coexisting species is at least as large
as the number of resource types. Such maximally diverse
communities typically require fine-tuning of the resource
utilization profiles or imposition of universal efficiency
tradeoffs in cellular metabolism.
In our stochastically assembled communities, the di-
versity is always much lower than the number of re-
source types, so we hypothesized that the resource con-
centrations should not be pinned. To test this idea, we
measured the response of the steady-state concentrations
¯Rα to changes in external supply rates κα, in terms of
the "resource susceptibilities" ∂ ¯Rα/∂κα plotted in Fig-
ure 4D [1]. Our hypothesis was valid in the resource-
limited regime, where many resource susceptibilities are
comparable to the susceptibility in the empty chemostat
∂ ¯Rα/∂κα = τR = 1. But in the diverse regime, we were
surprised to find that the susceptibilities are 100 times
smaller than this maximum value. This suggests that re-
6
FIG. 5. Richness of diverse regime depends on gen-
eralized niche-overlap. We took the values of supplied
energy flux w0κ0 and leakage fraction l from the three ex-
amples highlighted in Figure 2, and varied the average niche
overlap (cid:104)ρij(cid:105) between members of the metacommunity. For
each w0κ0, l combination and each value of (cid:104)ρij(cid:105), we gener-
ated 10 pools of 200 species, initialized 10 communities of 100
species each from this pool, and ran the dynamics to steady
state. The steady-state richness of each community is plotted
against the niche overlap. Points are colored by their regime
(diverse or resource-limited), and solid lines are linear regres-
sions.
Inset: ciα matrices that define the regional pool for
two different levels of overlap, with dark squares representing
high consumption coefficients.
regional pool is given by:
(cid:104)ρij(cid:105) ≡
(cid:42) (cid:80)
(cid:113)(cid:80)
(cid:80)
α ciαcjα
α c2
iα
(cid:43)
(cid:113)
=
(cid:104)ciα(cid:105)2
(cid:104)c2
iα(cid:105)(cid:104)c2
jα(cid:105)
.
(8)
α c2
jα
Figure 5 shows how the richness varies as a function of
(cid:104)ρij(cid:105). In the diverse regime the mean richness decreases
approximately linearly with increasing overlap. The rich-
ness of the resource-limited regime, on the other hand,
has only a very weak dependence on the niche overlap.
These results suggest that the distribution of consump-
tion preferences in the regional pool is the primary driver
of community assembly in the diverse regime.
Impor-
tantly, non-zero niche overlap limits the number of co-
existing species well below the upper bound imposed by
the competitive exclusion principle.
FIG. 4. Structure and stability of resource dynam-
ics depend on ecological regime. (A) Consumed energy
fluxes (1 − l)J in
iα for each of the ten surviving species in a
resource-limited community (example 2 from Figure 2). The
black portion of the bar is the flux (1 − l)J in
i0 due to the ex-
ternally supplied resource, and the colored bars represent the
contributions of the other resources. Since these communities
have reached the steady state, Equation (3) implies that the
total height of each bar equals the maintenance cost mi of the
corresponding consumer species. (B) Same as previous panel,
but for a community from the diverse regime (example 3 from
Figure 2). (C) Simpson diversity M eff
i of steady-state flux vec-
tor J in
iα for each species from examples 2 (resource-limited)
and 3 (diverse) in Figure 2. Vertical lines indicate the values
of this metric when all the flux is concentrated on a single
resource ("Specialist"), and where it is evenly spread over ten
resource types ("Generalist"). (D) Logarithm of susceptibil-
ity log10 ∂ ¯Rα/∂κα of community-supplied resources (α (cid:54)= 0)
to addition of an externally supplied flux κα in these two ex-
amples.
source pinning may be a generic phenomenon, observable
in real ecosystems when the energy supply is sufficiently
large.
Niche overlap limits richness in diverse regime
In the diverse regime, the number of coexisting species
("richness") is not limited by energy availability or by
access to secreted metabolites, but is still much less than
the maximal value of M = 100 set by the competitive ex-
clusion principle [32], even though almost all M resource
types are present at non-negligible levels (as shown in
Appendix Figure 18). We hypothesized that the diver-
sity in this regime is limited by the degree of similarity
between consumption preferences of members of the re-
gional species pool. This can be quantified in terms of
the niche overlap [10, 36], whose average value in a large
Nestedness and other large-scale beta-diversity
patterns
Our aim in developing this model is to identify and un-
derstand generic patterns in community structure, that
are independent of particular biological details. In large-
scale surveys of natural communities, subject to many
sources of noise and environmental heterogeneity, one
expects that only sufficiently generic patterns will be de-
tectable. The simplest observable to examine in such
survey data is the list of species that are present or ab-
sent in each sample. We obtained these presence/absence
(A)(B)(C)(D).7
only allow a few species from the regional pool to rise to
dominance, and that these dominant species induce clus-
tering of communities. Such "enterotype"-like behavior
is a common feature observed in many microbial settings
[3]. In contrast, the diverse regime exhibited neither well-
defined clusters nor dominant, highly abundant species.
Comparison to microbial datasets
The preceding results suggest that the resource-limited
and diverse regimes can be distinguished using beta-
diversity patterns. Rigorous testing of this prediction
is beyond the scope of the present work. But as an il-
lustration of the kind of data we hope to explain, we
examined the natural gradient of solar energy supply in
the Tara Oceans survey, which collected microbial com-
munity samples from a range of depths across the world's
oceans [47]. Explicitly including light as an energy source
would require some modification to the structure of the
model equations, but we expect that the large-scale fea-
tures of sufficiently diverse ecosystems should not be sen-
sitive to changes involving just one resource. We ana-
lyzed the 16S OTU composition of tropical ocean com-
munities for all 30 sea-surface samples, where solar en-
ergy is plentiful, and all 13 samples from the deep-sea
Mesopelagic Zone where energy fluxes are limited. We
projected these composition vectors onto their first two
principal components as in Figure 6 above, and plot the
results in Figure 7. The sea surface data superficially
resembles our diverse regime, with a relatively uniform
distribution of possible community compositions. In con-
trast, the Mesopelagic Zone is similar to our resource-
limited regime:
the dominance of the most abundant
species is much more pronounced, and the compositions
appear to cluster into four discrete types. While these
results are consistent with our model predictions, the
number of samples at each depth is still too small to
draw any definitive conclusions about clustering. As
mentioned above, our model also gives a natural explana-
tion for the nestedness in the Earth Microbiome Project
community composition data [49], suggesting that it may
be a natural byproduct of complex microbial communi-
ties shaped by competition, environmental heterogeneity,
and stochastic colonization. To test how generic this fea-
ture is, we plotted presence/absence community compo-
sitions of all samples from the Tara Oceans dataset, sort-
ing samples by richness and OTU's ("species") by preva-
lence. Each sample contains thousands of low-abundance
OTU's, which can obscure ecological patterns through
their susceptibility to sequencing noise and transient im-
migration. We therefore imposed a 0.5% relative abun-
dance threshold for an OTU to count as "present." The
resulting pattern in Figure 7 is qualitatively similar to
our simulations (Fig. 6D), and to the phylum-level data
of the Earth Microbiome Project [49], with the region
below the diagonal significantly less populated than the
region above the diagonal. In Appendix C 5 and the ac-
FIG. 6. Resource-limited regime features community-
level environmental filtering. (A) Presence (black) or ab-
sence (white) of all species in all 1,000 communities from the
original simulations of Figure 2. (B, C, D) We initialized 200
new communities for each of the three examples highlighted
in Figure 2A, by randomly choosing sets of 100 species from
the regional pool. Each panel shows the projection of final
community compositions {Ni} onto the first two principal
components of the set of compositions.
vectors from the simulations of Figure 2, and found that
when we sorted species by prevalence (rows in Fig. 6A)
and samples by richness (columns in Fig. 6A), the com-
munity composition generically exhibited a nested struc-
ture -- less diverse communities had a subset of the species
prevalent in the more diverse communities [33, 42]. We
quantified this result using an established nestedness
metric, as described below in Appendix C 5, and found
that the actual nestedness exceeds the mean value for a
randomized null model by more than 100 standard devi-
ations. This suggests that nested structures may gener-
ically emerge in community assembly through the inter-
play of stochastic colonization, competition, and environ-
mental filtering.
Next, we asked if we large-scale beta-diversity pat-
terns could be used to distinguish the resource-limited
and diverse regimes. We initialized 200 new communi-
ties with 100 randomly chosen members from the full
regional species pool and simulated these communities
to steady state in both the resource-limited and diverse
regimes (see Appendix B for details). This sub-sampling
of the full regional species pools mimics the effect of
stochastic colonization, where a different random sub-
set of species seeds each community. To better under-
stand beta-diversity signatures in the two regimes, we
performed a Principal Component Analysis (PCA) on
community composition and projected the data onto the
first two principal components, as shown in Figure 6B-
D. In the resource-limited regime, the communities form
distinct clusters that are distinguished by different highly
abundant species. This suggests that harsh environments
(A)(B)(C)(D)8
strongly constrained by species- and community-level en-
vironmental filtering. Each community is dominated by a
handful of species, making the community properties sen-
sitive to the idiosyncratic characteristics of these species
and susceptible to environmental fluctuations. In the di-
verse phase, communities exhibit more universal features
because they substantially engineer their environments.
In particular, the concentrations of resources at steady
state are more narrowly distributed and insensitive to
perturbations in the external supply rates. Moreover,
each species draws its energy roughly equally from all re-
sources, rather than subsisting on the externally supplied
resource as in the resource-limited phase.
The emergence of environmental engineering from this
community-scale model makes it a valuable tool for test-
ing and refining existing conceptual
frameworks em-
ployed by empirical biologists [46]. A major limitation
of the dominant paradigms for evolution and ecology
from the last century is the implicit assumption of a con-
stant environment [14]. The generalized Lotka-Volterra
model, for example, remains a standard lens for reason-
ing about ecological dynamics, both quantitatively and
qualitatively [17, 20, 54].
It assumes that the dynam-
ics emerge from the sum of pairwise interactions among
species, and that the sign and strength of these interac-
tions are intrinsic properties of the species. This can be
a good assumption in some circumstances [20, 54], but
fails to accurately describe the behavior of simple models
that explicitly account for the state of the environment
[39]. Our work provides a starting point for determining
the conditions under which pairwise models will generi-
cally succeed or fail in describing the behavior of large
ecosystems.
Our model complements other recent efforts at under-
standing microbial community ecology. Taillefumier et
al. proposed a similar model of metabolite exchange,
and focused on the case where the number of resource
types M is equal to 3 [48]. In this case, repeated inva-
sion attempts from a large regional species pool produced
optimal combinations of metabolic strategies. Goyal et
al. examined the opposite limit, with M = 5, 000, but
allowed each species to consume only one type of re-
source [25]. This generated communities with a tree-like
metabolic structure, where each species depends directly
on another species to generate its unique food source. In
our model, the large number of resource types (M = 100
in the current study) makes spontaneous strategy opti-
mization extremely unlikely. And our generic protocol
for sampling the metabolic matrix Dαβ allows a variety
of community-level energy flux topologies to emerge, as
illustrated in Figure 3, which can sometimes be quite
different from the tree networks of Goyal et al.. The
absence of highly specialized metabolic structure in our
model makes it especially well-suited for interpreting pat-
terns in large-scale sequence-based datasets such as the
Earth Microbiome Project [49].
Our model predictions can also be directly tested using
experiments with natural communities in synthetic lab-
Ecological regimes and nestedness in mi-
FIG. 7.
(A) 16S OTU compositions of tropical
crobiome data.
mesopelagic zone samples from the Tara Oceans database,
collected at a depth of 200 to 1,000 meters [47]. Each dot is
the projection of one sample onto the first two principal com-
ponents of the collection of mesopelagic zone samples. (B)
Same as A, but for tropical surface water layer samples, col-
lected at a depth of 5 meters. (C) Presence (black) or absence
(white) of each OTU above 0.5% relative abundance across
all Tara Oceans samples.
companying Figure 14, we quantify the nestedness us-
ing the same metric employed in the Earth Microbiome
Project analysis [2, 49], and show that the score is signif-
icantly higher than the mean scores from two standard
null models.
DISCUSSION
Advances in sequencing technology have opened new
horizons for the study of microbial ecology, generating
massive amounts of data on the composition of both nat-
ural and synthetic communities. But the complexity of
these systems make it difficult to extract robust general
principles suitable for guiding medical and industrial ap-
plications. Numerical simulations provide a powerful tool
for addressing this problem. By rapidly iterating numer-
ical experiments under a variety of modeling choices with
random parameters, one can identify robust patterns and
use the resulting insights to guide targeted experiments.
Following this strategy, we developed a thermody-
namic consumer resource model that explicitly includes
energetic fluxes and metabolically-mediated cross-feeding
and competition. Using this model, we identified two
qualitatively distinct phases in these simulations as we
varied the amount of energy supplied to ecosystem and
the fraction of energy leaked back into the environment:
a low diversity "resource-limited" phase and a "diverse"
phase. The structure of the resource-limited phase is
(A)(C)(B)9
oratory environments [12, 23]. Our model predicts that
beta-diversity patterns and community-level metabolic
networks can be significantly altered by increasing the
ecosystem's energy supply, inducing a transition from the
resource-limited to the diverse regime. In the experimen-
tal set-up of [12], this can be done by directly adding
chitinase enzymes to the sludge reactor to increase the
degradation of chitin-based organic particles on which
the ocean-derived microbial communities subsist. One
could then look for shifts in the resulting diversity pat-
terns, and observe any changes in the topology of the
metabolic flux network using isotope labeling.
In this work we have largely confined ourselves to
studying steady-state properties of well-mixed microbial
communities. Microbial communities often exhibit com-
plex temporal dynamics with well-defined successions
[12, 16, 57]. It will be interesting to explore these dynami-
cal phenomena using our model. It is also well established
that spatial structure can give rise to new ecological phe-
nomena [31, 38] and an important area of future work
will be to better explore the role of space in microbial
community assembly.
Finally, we have obtained strong numerical evidence
that the two regimes are separated by a phase transition,
which is likely closely related to disorder-induced phase
transitions in statistical physics [6]. In Appendix D, we
examine the steady-state richness in the three examples
of Figure 2 under increasing values of M from M = 40
to M = 560. We find that the richness is proportional to
M in the diverse regime, but scales sub-linearly with M
in both examples from the resource-limited regime.
In
the M → ∞ limit, therefore, we expect to find a sharp
line between the regimes, with the ratio of the richness
to M vanishing on the resource-limited side. But we do
not yet know the exact location of this boundary, or the
critical exponents describing the behavior of the system
near the transition.
ACKNOWLEDGMENTS
The funding for this work partly results from a Scialog
Program sponsored jointly by Research Corporation for
Science Advancement (RCSA) and the Gordon and Betty
Moore Foundation. This work was also supported by
NIH NIGMS grant 1R35GM119461, by a Cottrell Scholar
award from RCSA to KK, and by Simons Investigator in
the Mathematical Modeling of Living Systems (MMLS)
awards to PM and KK. The authors are pleased to ac-
knowledge that the computational work reported on in
this paper was performed on the Shared Computing Clus-
ter which is administered by Boston Universitys Research
Computing Services.
[1] Madhu Advani, Guy Bunin, and Pankaj Mehta. Statis-
tical physics of community ecology: a cavity solution to
MacArthur's consumer resource model. Journal of Sta-
tistical Mechanics, page 033406, 2018.
[2] M´ario Almeida-Neto, Paulo Guimaraes, Paulo R
Guimaraes Jr, Rafael D Loyola, and Werner Ulrich. A
consistent metric for nestedness analysis in ecological
systems: reconciling concept and measurement. Oikos,
117(8):1227, 2008.
[3] Manimozhiyan Arumugam, Jeroen Raes, Eric Pelletier,
Denis Le Paslier, Takuji Yamada, Daniel R Mende,
Gabriel R Fernandes, Julien Tap, Thomas Bruls, Jean-
Michel Batto, et al. Enterotypes of the human gut mi-
crobiome. Nature, 473:174, 2011.
[4] Matthieu Barbier, Jean-Fran¸cois Arnoldi, Guy Bunin,
and Michel Loreau. Generic assembly patterns in com-
plex ecological communities. Proceedings of the National
Academy of Sciences, 2018.
[5] A. Bekker, H. D. Holland, P.-L. Wang, D. Rumble III,
H. J. Stein, J. L. Hannah, L. L. Coetzee, and N. J.
Beukes. Dating the rise of atmospheric oxygen. Nature,
427:117, 2004.
[6] Giulio Biroli, Guy Bunin, and C. Cammarota. Marginally
stable equilibria in critical ecosystems. New Journal of
Physics, 20:083051, 2018.
[7] G. Bunin. Ecological communities with Lotka-Volterra
dynamics. Physical Review E, 95:042414, 2017.
[8] Stacey Butler and James O'Dwyer. Stability criteria for
complex microbial communities. bioRxiv, 2018.
[9] Jonathan M Chase. Community assembly: when should
history matter? Oecologia, 136:489, 2003.
[10] Peter Chesson. Macarthur's consumer-resource model.
Theoretical Population Biology, 37:26, 1990.
[11] Luca D'Alessio, Yariv Kafri, Anatoli Polkovnikov, and
Marcos Rigol. From quantum chaos and eigenstate ther-
malization to statistical mechanics and thermodynamics.
Advances in Physics, 65:239, 2016.
[12] Manoshi S. Datta, Elzbieta Sliwerska, Jeff Gore, Mar-
tin F. Polz, and Otto X. Cordero. Microbial interactions
lead to rapid micro-scale successions on model marine
particles. Nature Communications, 7:11965, 2016.
[13] Benjamin Dickens, Charles K. Fisher, and Pankaj Mehta.
Analytically tractable model for community ecology with
many species. Physical Review E, 94:022423, 2016.
[14] Michael Doebeli, Yaroslav Ispolatov, and Burt Simon.
Point of view: Towards a mechanistic foundation of evo-
lutionary theory. Elife, 6:e23804, 2017.
[15] Mallory Embree, Joanne K Liu, Mahmoud M Al-
Bassam, and Karsten Zengler. Networks of energetic
and metabolic interactions define dynamics in microbial
communities. Proceedings of the National Academy of
Sciences, 112:15450, 2015.
[16] Tim N. Enke, Gabriel E. Leventhal, Matthew Metzger,
Jos´e T. Saavedra, and Otto X. Cordero. Micro-scale
ecology regulates particulate organic matter turnover in
model marine microbial communities. bioRxiv, 2018.
[17] Farnoush Farahpour, Mohammadkarim Saeedghalati,
Verena S Brauer, and Daniel Hoffmann. Trade-off shapes
diversity in eco-evolutionary dynamics. eLife, 7:e36273,
2018.
[18] Charles K. Fisher and Pankaj Mehta. The transition
between the niche and neutral regimes in ecology. PNAS,
111:13111, 2014.
[19] Jonathan Friedman, Logan M. Higgins, and Jeff Gore.
Community structure follows simple assembly rules in
microbial microcosms. Nature Ecology and Evolution,
1:0109, 2017.
[20] Jonathan Friedman, Logan M Higgins, and Jeff Gore.
Community structure follows simple assembly rules in
microbial microcosms.
Nature ecology & evolution,
1:0109, 2017.
[21] G. F. Gause and A. A. Witt. Behavior of mixed popula-
tions and the problem of natural selection. The American
Naturalist, 69:596, 1935.
[22] Theo Gibbs, Jacopo Grilli, Tim Rogers, and Stefano
Allesina. The effect of population abundances on the
stability of large random ecosystems. arXiv, 1708.08837,
2017.
[23] Joshua E. Goldford, Nanxi Lu, Djordje Baji´c, Sylvie
Estrela, Mikhail Tikhonov, Alicia Sanchez-Gorostiaga,
Daniel Segr`e, Pankaj Mehta, and Alvaro Sanchez. Emer-
gent simplicity in microbial community assembly. Sci-
ence, 361:469, 2018.
[24] Benjamin H. Good, Stephen Martis, and Oskar Hal-
latschek. Directional selection limits ecological diversifi-
cation and promotes ecological tinkering during the com-
petition for substitutable resources. bioRxiv, 2018.
[25] Akshit Goyal and Sergei Maslov. Diversity, stability,
and reproducibility in stochastically assembled microbial
ecosystems. Physical Review Letters, 120:158102, 2018.
[26] William R. Harcombe, William J. Riehl, Ilija Dukovski,
Brian R. Granger, Alex Betts, Alex H. Lang, Gracia
Bonilla, Amrita Kar, Nicholas Leiby, Pankaj Mehta,
Christopher J. Marx, and Daniel Segr`e. Metabolic
resource allocation in individual microbes determines
ecosystem interactions and spatial dynamics. Cell Re-
ports, 7:1104, 2014.
[27] Curtis Huttenhower, Dirk Gevers, Rob Knight, Sa-
har Abubucker, Jonathan H Badger, Asif T Chinwalla,
Heather H Creasy, Ashlee M Earl, Michael G FitzGerald,
Robert S Fulton, et al. Structure, function and diversity
of the healthy human microbiome. Nature, 486:207, 2012.
[28] Patricio Jeraldo, Maksim Sipos, Nicholas Chia, Jen-
nifer M Brulc, A Singh Dhillon, Michael E Konkel,
Charles L Larson, Karen E Nelson, Ani Qu, Lawrence B
Schook, F Yang, Bryan A White, and Nigel Golden-
feld. Quantification of the relative roles of niche and
neutral processes in structuring gastrointestinal micro-
biomes. Proceedings of the National Academy of Sciences,
109:9692, 2012.
[29] Clive G. Jones, John H. Lawton, and Shachak Moshe.
Organisms as ecosystem engineers. Oikos, 69:373, 1994.
[30] David A. Kessler and Nadav M. Shnerb. General-
ized model of island biodiversity. Physical Review E,
91:042705, 2015.
[31] Kirill S Korolev, Mikkel Avlund, Oskar Hallatschek, and
David R Nelson. Genetic demixing and evolution in lin-
ear stepping stone models. Reviews of Modern Physics,
82:1691, 2010.
[32] Simon A Levin. Community equilibria and stability, and
an extension of the competitive exclusion principle. The
American Naturalist, 104:413, 1970.
10
[33] Mark V. Lomolino. Investigating causality of nestedness
of insular communities: Selective immigrations or extinc-
tions? Journal of Biogeography, 23:699, 1996.
[34] Michel Loreau. Consumers as maximizers of matter and
energy flow in ecosystems. The American Naturalist,
145:22, 1995.
[35] Robert MacArthur. Species Packing and Competitive
Equilibrium for Many Species. Theoretical Population
Biology, 1:1, 1970.
[36] Robert MacArthur and Richard Levins. The Limiting
Similarity, Convergence, and Divergence of Coexisting
Species. The American Naturalist, 101:377, 1967.
[37] Robert M. May. Stability and complexity in model ecosys-
tems. Princeton University Press, Princeton, N.J., 2001.
[38] Rajita Menon and Kirill S Korolev. Public good diffu-
sion limits microbial mutualism. Physical Review Letters,
114:168102, 2015.
[39] Babak Momeni, Li Xie, and Wenying Shou. Lotka-
volterra pairwise modeling fails to capture diverse pair-
wise microbial interactions. Elife, 6:e25051, 2017.
[40] Hidetoshi Nishimori. Statistical Physics of Spin Glasses
and Information Processing. Oxford University Press,
New York, NY, 2001.
[41] Benjamin Obadia, Z.T. Guvener, Vivian Zhang,
Javier A. Ceja-Navarro, Eoin L. Brodie, William W. Ja,
and William B. Ludington. Probabilistic invasion under-
lies natural gut microbiome stability. Current Biology,
27:1999, 2017.
[42] Bruce D Patterson and Wirt Atmar. Nested subsets and
the structure of insular mammalian faunas and archipela-
gos. Biological Journal of the Linnean Society, 28:65,
1986.
[43] F. Pedregosa, G. Varoquaux, A. Gramfort, V. Michel,
B. Thirion, O. Grisel, M. Blondel, P. Prettenhofer,
R. Weiss, V. Dubourg, J. Vanderplas, A. Passos, D. Cour-
napeau, M. Brucher, M. Perrot, and E. Duchesnay.
Scikit-learn: Machine learning in Python. Journal of Ma-
chine Learning Research, 12:2825 -- 2830, 2011.
[44] Anna Posfai, Thibaud Taillefumier,
and Ned S.
Wingreen. Metabolic trade-offs promote diversity in a
model ecosystem. Physical Review Letters, 118:028103,
2017.
[45] Bettina E. Schirrmeister, Muriel Gugger, and Philip C. J.
Donoghue. Cyanobacteria and the great oxidation event:
Evidence from genes and fossils. Palaeontology, 58:769,
2015.
[46] Wenying Shou, Carl T Bergstrom, Arup K Chakraborty,
and Frances K Skinner. Theory, models and biology.
Elife, 4:e07158, 2015.
[47] Shinichi Sunagawa, Luis Pedro Coelho, Samuel Chaf-
fron, Jens Roat Kultima, Karine Labadie, Guillem
Salazar, Bardya Djahanschiri, Georg Zeller, Daniel R.
Mende, Adriana Alberti, Francisco M. Cornejo-Castillo,
Paul I. Costea, Corinne Cruaud, Francesco d'Ovidio,
Stefan Engelen, Isabel Ferrera, Josep M. Gasol, Lionel
Guidi, Falk Hildebrand, Florian Kokoszka, Cyrille Lep-
oivre, Gipsi Lima-Mendez, Julie Poulain, Bonnie T. Pou-
los, Marta Royo-Llonch, Hugo Sarmento, Sara Vieira-
Silva, C´eline Dimier, Marc Picheral, Sarah Searson, Ste-
fanie Kandels-Lewis, Chris Bowler, Colomban de Var-
gas, Gabriel Gorsky, Nigel Grimsley, Pascal Hingamp,
Daniele Iudicone, Olivier Jaillon, Fabrice Not, Hiroyuki
Ogata, Stephane Pesant, Sabrina Speich, Lars Stem-
mann, Matthew B. Sullivan, Jean Weissenbach, Patrick
Wincker, Eric Karsenti, Jeroen Raes, Silvia G. Acinas,
and Peer Bork. Structure and function of the global ocean
microbiome. Science, 348:1261359, 2015.
[48] Thibaud Taillefumier, Anna Posfai, Yigal Meir, and
Ned S. Wingreen. Microbial consortia at steady supply.
eLife, 6:e22644, 2017.
[49] Luke R. Thompson, Jon G. Sanders, Daniel McDon-
ald, Amnon Amir, Joshua Ladau, Kenneth J. Locey,
Robert J. Prill, Anupriya Tripathi, Sean M. Gibbons,
Gail Ackermann, Jose A. Navas-Molina, Stefan Janssen,
Evguenia Kopylova, Yoshiki V´azquez-Baeza, Antonio
Gonz´alez, James T. Morton, Siavash Mirarab, Zhen-
jiang Zech Xu, Lingjing Jiang, Mohamed F. Haroon,
Jad Kanbar, Qiyun Zhu, Se Jin Song, Tomasz Kosciolek,
Nicholas A. Bokulich, Joshua Lefler, Colin J. Brislawn,
Gregory Humphrey, Sarah M. Owens, Jarrad Hampton-
Marcell, Donna Berg-Lyons, Valerie McKenzie, Noah
Fierer, Jed A. Fuhrman, Aaron Clauset, Rick L. Stevens,
Ashley Shade, Katherine S. Pollard, Kelly D. Goodwin,
Janet K. Jansson, Jack A. Gilbert, Rob Knight, and
Earth Microbiome Project Consortium. A communal
catalogue reveals Earth's multiscale microbial diversity.
Nature, 551:457, 2017.
[50] Mikhail Tikhonov and Remi Monasson. Collective phase
in resource competition in a highly diverse ecosystem.
Physical Review Letters, 118:048103, 2017.
[51] Mikhail Tikhonov and Remi Monasson.
Innovation
Rather than Improvement: A Solvable High-Dimensional
Model Highlights the Limitations of Scalar Fitness. Jour-
nal of Statistical Physics, 2018.
[52] Hanna Tuomisto. A consistent terminology for quanti-
fying species diversity? Yes, it does exist. Oecologia,
164:853, 2010.
[53] Nicole M. Vega and Jeff Gore. Stochastic assembly pro-
duces heterogeneous communities in the Caenorhabditis
elegans intestine. PLoS Biol., 15:e2000633, 2017.
[54] Ophelia S Venturelli, Alex V Carr, Garth Fisher, Ryan H
Hsu, Rebecca Lau, Benjamin P Bowen, Susan Hromada,
Trent Northen, and Adam P Arkin. Deciphering mi-
crobial interactions in synthetic human gut microbiome
communities. Molecular systems biology, 14:e8157, 2018.
[55] Stefanie Widder, Rosalind J Allen, Thomas Pfeiffer,
Thomas P Curtis, Carsten Wiuf, William T Sloan,
Otto X Cordero, Sam P Brown, Babak Momeni, Weny-
ing Shou, Helen Kettle, Harry J Flint, Andreas F Haas,
B´eatrice Laroche, Jan-Ulrich Kreft, Tobias Grosskopf, Jef
Huisman, Andrew Free, Cristian Picioreanu, Christopher
Quince, Isaac Klapper, Simon Labarthe, Barth F. Smets,
Harris Wang, and Orkun S Soyer. Challenges in microbial
ecology: building predictive understanding of commu-
nity function and dynamics. The ISME journal, 10:2557,
2016.
[56] Eugene P. Wigner. Characteristic vectors of bordered
matrices with infinite dimensions. Annals of Mathematics
(ser. 2), 62:548, 1955.
[57] Benjamin E. Wolfe and Rachel J. Dutton. Fermented
foods as experimentally tractable microbial ecosystems.
Cell, 161:49, 2015.
[58] Ali R. Zomorrodi and Daniel Segr`e. Synthetic ecology
of microbes: Mathematical models and applications. J.
Mol. Biol., 428:837, 2016.
11
SI Appendices
In Appendix A, we provide a full explanation of the
Microbial Consumer Resource Model. We describe
its numerical
implementation in Appendix B, and
present additional data illustrating the robustness of the
qualitative results in Appendix C. Finally, Appendix
D contains preliminary evidence for the existence of a
bona fide phase transition in the M → ∞ limit. All data
and code for generating figures can be found at https:
//github.com/Emergent-Behaviors-in-Biology/
crossfeeding-transition.
Appendix A: Model details
1. Generalities
We begin by defining an energy flux into a cell J in,
an energy flux that is used for growth J growth, and an
outgoing energy flux due to byproduct secretion J out.
Energy conservation requires
J in = J growth + J out
(A1)
for any reasonable metabolic model. Now consider a
model with M resources Rβ with β = 1 . . . M each with
"energy" or quality wβ. It will be useful to divide the in-
put and output energy fluxes that are consumed/secreted
in metabolite β by J in
respectively. We define
the fraction f out
of the output energy secreted as resource
β by
β and J out
β
β
J out
β ≡ f out
β J out.
We can define corresponding mass fluxes by
and
νout
β ≡ J out
β /wβ
νin
β ≡ J in
β /wβ
(A2)
(A3)
(A4)
In general, all these fluxes depend on the species under
consideration and will carry an extra roman index i in-
dicating the species.
We assume that a fixed quantity mi of power per cell
is required for maintenance of species i, and that the
per-capita growth rate is proportional to the remaining
energy flux (J growth − mi), with proportionality constant
gi. Under these assumptions, the time-evolution of the
population size Ni of species i can be modeled using the
equation
dNi
dt
= giNi(J growth
i
− mi).
(A5)
We can model the resource dynamics by functions of the
form
dRα
dt
= hα(Rα) −
Njνin
jα +
Njνout
jα ,
(A6)
j
j
(cid:88)
(cid:88)
12
A direct consequence of energy conservation (Equation
(A1)) is that
(cid:88)
(1 − lα)J in
α
iα =
(cid:88)
α
J growth
i
=
(1 − lα)wασ(ciαRα)
α. Note that by definition(cid:80)
(A14)
All that is left is to determine how to compute the prob-
ability of producing a byproduct β when consuming α.
Let us denote by Dβα the fraction of the output energy
that is contained in metabolite β when a cell consumes
β Dβα = 1. These Dβα and
lα uniquely specify the metabolic model for independent
resources and we can write all fluxes in terms of these
quantities.
The total energy output in metabolite β is thus
(cid:88)
(cid:88)
J out
iβ =
DβαlαJ in
iα =
Dβαlαwασ(ciαRα).
(A15)
α
This also yields
νout
iβ =
α
Dβαlα
wα
wβ
(cid:88)
α
σ(ciαRα)
(A16)
We are now in position to write down the full dynamics
in terms of these quantities:
(cid:35)
(cid:34)(cid:88)
(cid:88)
j
α
(cid:88)
dNi
dt
dRα
dt
= giNi
(1 − lα)wασ(ciαRα) − mi
= hα(Rα) −
Njσ(cjαRα)
(cid:20)
(cid:21)
wβ
wα
lβ
(A17)
+
Njσ(cjβRβ)
Dαβ
where the function hα describes the resource dynamics
in the absence of consumers. We can consider two kinds
of dynamics: externally supplied and self-renewing. For
externally supplied resources, we take a linearized form
of the dynamics:
hexternal
α
(Rα) = κα − τ−1
α Rα
(A7)
while for self-renewing we take a logistic form for the
dynamics
hself−renewing
α
(Rα) = rαRα(Kα − Rα).
(A8)
In the present study, we only consider externally supplied
resources.
These equations specify the general dynamics of all
the models we consider. Metabolism is encoded in the
relationship between input, output, and growth fluxes.
2.
Input fluxes and output partitioning
We will now specify the form of the input fluxes νin
β
and the output partitioning f out
β . This involves specify-
ing how an input resource is turned into an metabolic
byproducts. To try to capture metabolic structure, we
will divide the M resources into T classes (e.g.
sug-
ars, amino acids, etc.), each with MA resources where
A MA = M . We will be interested
in capturing coarse metabolic structure (i.e. metaboliz-
ing sugars outputs carboxylic acids, etc). We will limit
ourselves to considering strictly substitutable resources.
A = 1, . . . T and (cid:80)
In all consumer resource models, we assume that
νin
iβ = σ(ciβRβ)
(A9)
jβ
where σ encodes the response function of consumer i for
resource α. In the microbial context the consumer prefer-
ences ciα can be interpreted as expression levels of trans-
porters for each of the resources. We consider three kinds
of response functions: Type-I, linear response functions
where
σI (x) = x,
a Type-II saturating Monod function,
σII (x) =
x
1 + x
K
and a Type-III Hill or sigmoid-like function
σIII (x) =
1 +(cid:0) x
xn
K
(cid:1)n ,
(A10)
(A11)
(A12)
where n > 1.
In all the simulations of this paper, we assume that
resources independently contribute to the growth rate.
We define a leakage fraction 0 ≤ lα ≤ 1 for resource α
such that
α = lαJ in
J out
α .
(A13)
Notice that when σ is Type-I (linear) and lα = 0 for all α
(no leakage or byproducts), this reduces to MacArthur's
original model [35].
3. Choosing consumer preferences
We will now choose consumer preferences ciα as fol-
lows. We assume that each specialist family has a pref-
erence for one resource class A (where A = 1 . . . F ) with
0 ≤ F ≤ T , and we denote the consumer coefficients for
this family by cA
iα. We will also consider generalists that
have no preferences, with consumer coefficients cgen
iα . We
will consider three kinds of models: one where the coef-
ficients are drawn from Gaussian distributions, another
where they are drawn from Gamma distributions (which
ensure positivity of the coefficients), and finally a dis-
crete, binary preference model.
a. Gaussian consumer preferences
The Gaussian model allows a continuous gradation of
transporter expression levels. We assume that the vari-
ance is fixed to so that for all coefficients for all families
for the generalists.
13
(cid:104)(δcA
iα)2(cid:105) = (cid:104)(δcgen
iα )2(cid:105) =
σ2
c
M
.
(A18)
In the generalist family, the mean is also the same for all
resources, and is given by
Note that the mean of the distribution is (cid:104)ciα(cid:105) =
piαc1 + c0
M , and the variance is (cid:104)(δciα)2(cid:105) = piα(1 − piα).
Both of these scale as 1/M when M → ∞, just like the
Gaussian and Gamma versions, as long as c1, c0 and µc
are held fixed.
(cid:104)cgen
iα (cid:105) =
µc
M
.
(A19)
4. Constructing the metabolic matrix
(cid:104)
(cid:40) µc
M
µc
(cid:105)
The specialist families sample from a distribution with a
larger mean for resources in their preferred class:
(cid:104)cA
iα(cid:105) =
1 + M−MA
MA
qA
M (1 − qA),
,
if α ∈ A
otherwise,
(A20)
where MA is the number of resources in class A, and
qA controls how much more species from family A prefer
resources from class A.
We have put a factor of M in the denominators of
the expressions for mean and variance, because the sums
over ciα in the dynamical equations (A17) always give
rise to terms with means M(cid:104)ciα(cid:105) or S(cid:104)ciα(cid:105) and variances
M(cid:104)(δciα)2(cid:105) or S(cid:104)(δciα)2(cid:105). The factor of M allows us to
keep σc, µc fixed when exploring the M, S → ∞ limit in
Appendix D below.
b. Gamma consumer preferences
We will also consider the case where consumer prefer-
ences are drawn from Gamma distributions, which guar-
antee that all coefficients will be positive. Since the
Gamma distribution only has two parameters, it is fully
determined once the mean and variance are specified. We
parameterize the mean and variance for this model in the
same way as for the Gaussian model.
c. Binary consumer preferences
In the binary model, there are only two possible ex-
M and a
pression levels for each transporter: a low level c0
high level c0
M + c1. The elements of cA
iα are given by
cA
iα =
c0
M
+ c1Xiα,
(A21)
where Xiα is a binary random variable that equals 1 with
probability
(cid:40) µc
(cid:104)
(cid:105)
pA
iα =
M c1
µc
M c1
qA
MA
1 + M−MA
(1 − qA),
,
if α ∈ A
otherwise
(A22)
for the specialist families, and
pgen
iα =
µc
M c1
(A23)
We choose the metabolic matrix Dαβ according to a
three-tiered secretion model. The first tier contains a
preferred class of byproducts, such as carboyxlic acids
for fermentative and respiro-fermentative bacteria, which
includes Mc members. The second contains byproducts
of the same class as the input resource (when the import
resource is not in the preferred byproduct class). For ex-
ample, this could be attributed to the partial oxidation
of sugars into sugar alcohols, or the antiporter behavior
of various amino acid transporters. The third tier in-
cludes everything else. We encode this structure in Dαβ
by sampling each column of the matrix from a Dirichlet
distribution with concentration parameters dαβ that de-
pend on the byproduct tier, so that on average a fraction
fc of the secreted flux goes to the first tier, while a frac-
tion fs goes to the second tier, and the rest goes to the
third:
d0
d0
d0
d0
fc+fs
,
Mc
1−fc−fs
M−Mc
fs
,
MA(β)
,
1−fs−fc
M−MA(β)−Mc
,
if α = c
if α (cid:54)= c and β = c
if α, β (cid:54)= c and A(α) = A(β)
if α, β (cid:54)= c and A(α) (cid:54)= A(β).
(A24)
dαβ =
The parameter d0 controls the randomness of the parti-
tioning, ranging from the maximum value where all the
weight is put on a single resource for d0 = 1, to deter-
ministic partitioning as d0 → ∞.
The mean of the Dirichlet distribution is always equal
to 1/M , and the variance under this parameterization
also scales as 1/M when the f 's and d0 are held fixed.
The sampling of Dαβ thus following the same scaling be-
havior as our scheme for the consumer matrices in the
M, S → ∞ limit of Appendix D.
Appendix B: Simulation and data analysis
1. The Community Simulator
We implemented the above modeling framework
in a Python package called "Community Simulator,"
which can be downloaded and installed from https:
//github.com/Emergent-Behaviors-in-Biology/
community-simulator.
Once this package is
stalled,
//github.com/Emergent-Behaviors-in-Biology/
in-
the data can be downloaded from https:
crossfeeding-transition,
and the accompanying
Jupyter notebook can be used to regenerate all the
figures. The one exception is the energy flux network
figure, which was generated in MATLAB using a file ex-
ported from the notebook. The repository also contains
a sample MATLAB script for loading and visualizing
the network file.
Community Simulator is designed to run dynamics on
multiple communities in parallel, inspired by the paral-
lel experiments commonly performed with 96-well plates.
The central object of the package is a Community class,
whose instances are initialized by specifying the initial
population sizes and resource concentrations for each par-
allel "well," along with the functions and parameters that
define the population dynamics. This class contains two
core methods. Propagate(T) sends each community to
a separate CPU (for however many CPU's are available),
runs the given population dynamics for a time T using
the SciPy function odeint, and updates the population
sizes and resource concentrations in each well to the time-
evolved values. Passage(f) initializes a fresh set of wells
by adding a fraction fµν of the contents of each old well
ν to each new well µ. (Fresh media can also be added
at this point, but this feature was not relevant for the
current work). The resulting values of Ni are converted
from arbitrary concentration units to actual population
sizes using a specified scale factor, and then integer pop-
ulation sizes are obtained by multinomial sampling based
on these values.
The Community Simulator package also contains a
set of scripts for generating models and randomly
sampling parameters.
MakeConsumerDynamics and
MakeResourceDynamics from the usertools module
take a dictionary of assumptions concerning the re-
sponse type, metabolic regulation and resource replen-
ishment (as described above), and generate the corre-
sponding functions for dNi/dt and dRα/dt. The func-
tion MakeMatrices, from the same module, samples the
consumer matrix ciα and the metabolic matrix Dαβ ac-
cording to the scheme described in the previous section.
2. Simulation Details
For this paper, we generated a binary consumer matrix
with c0 = 1, c1 = 1 and µc = 10, and a metabolic matrix
with d0 = 0.2. This matrix defined a regional pool of
S = 200 species, consuming M = 100 possible resource
types. We used only one family and one resource class in
constructing the ciα and Dαβ matrices (but arbitrarily
assigned each resource and each species to one of four
categories, as a null model for comparison with future
structured simulations). We set wα = gi = 1 for all i and
α, and set the lα for all resources equal to each other. For
the multinomial sampling described above, we chose the
scale factor so that Ni = 1 corresponds to a population of
106 cells. We generated dynamics with Type-I response,
no regulation, and a "renewable" resource replenishment
14
model. Only resource type 0 was supplied externally,
with flux κ0, and all the other κα's were set to zero.
To simulate stochastic colonization, we initialized each
of 10 wells with 100 randomly chosen species from the
regional pool, with a population size of 106 cells per
species per well. We propagated each well under Equa-
tions (A17) for a time ∆t = 11, 500, which is much longer
than the maximum time required to relax to the steady
state for any of the parameter regimes sampled. We used
the Passage method with fµν = δµν to periodically elim-
inate species whose populations became too small. For
the large steady-state population sizes we consider here
(∼ 104− 109, see Figure 11 below), the multinomial sam-
pling eliminates species whose populations are heading
for extinction while minimally perturbing the dynamics
of the survivors. We passaged after every 5 time units
of propagation from the beginning of the simulation up
to time t = 500, then every 100 time units until time
t = 1, 500, and finally every 1,000 time units up to the
final time t = 11, 500.
The timeseries shown in Figure 1E was generated un-
der these assumptions, with w0κ0 = 500.
We propagated these 10 initial states using this proce-
dure for 100 different combinations of externally supplied
energy flux w0κ0 and leakage fraction l, with 10 w0κ0 val-
ues evenly spaced on a logarithmic scale from 10 to 100,
and 10 l values evenly spaced from 0 to 0.9. Figure 2
of the main text shows the mean richness over the 10
parallel wells for each combination of w0κ0 and l. The
richness is defined as the number of species with non-zero
abundance at the end of the simulation.
We focused on three representative examples for fur-
ther analysis:
1. Syntrophy-Limited: w0κ0 = 1000, (cid:104)lα(cid:105) = 0.1
2. Energy-Limited: w0κ0 = 28, (cid:104)lα(cid:105) = 0.6
3. Similarity-Limited: w0κ0 = 1000, (cid:104)lα(cid:105) = 0.9.
(cid:80)
The rank-abundance plots in Figure 2 of the main text
show the population sizes in all 10 wells from each of these
examples, after normalizing them by the total biomass
i Ni. The plots were truncated at a relative abundance
of 0.5% for clarity. Rank-abundance plots for these same
three examples in absolute units with no truncation can
be found in Figure 11 below.
3. Susceptibilities
One important property of an ecosystem is its sensi-
tivity to changes in environmental conditions. Figure 3
of the main text quantifies this sensitivity in terms of a
set of susceptibilities, defined by
∂ ¯Rα
∂κβ
∂ ¯Ni
∂κβ
χαβ ≡
ηiβ ≡
(B1)
(B2)
where ¯Ni, ¯Rα are the steady-state consumer populations
and resource concentrations, respectively.
5. Beta Diversity
15
For the case of externally supplied resources and Type-
I growth, setting Equations (A17) equal to zero and dif-
ferentiating with respect to κβ yields:
0 =
−τ−1
α δαβ =
(1 − lα)wαciαχαβ
cjγNj
Dαγ
(cid:20)
(cid:20)
×χγβ +
cjγ
Dαγ
(cid:88)
(cid:88)
α
γ
j
(cid:88)
(cid:88)
(cid:18)(cid:88)
(cid:19)
(cid:20)
(cid:88)
α δγα
jγ
γ
j
(cid:88)
− τ−1
+
jγ
(cid:21)
(cid:21)
lγ − δγα
lγ − δγα
wγ
wα
wγ
wα
(B3)
− τ−1
α δγα
Rγηjβ (B4)
(cid:20)
(cid:21)
The last equation can be reorganized as
−τ−1
α δαβ =
cjγNj
Dαγ
wγ
wα
lγ − δγα
χγβ
(cid:21)
wγ
wα
lγ − δγα
Rγηjβ
(B5)
cjγ
Dαγ
For each value of β, this system of linear equations can
be solved for χγβ and ηjβ by simply inverting a matrix
(once the terms corresponding to extinct species have
been removed).
The histograms of Figure 3D in the main text contain
the diagonal elements χαα for all resources except for
the one supplied externally (α = 0), which might be ex-
pected to behave somewhat differently. The χαα values
from all 10 parallel communities are included in the his-
togram. We generated one histogram for the similarity-
limited regime, and one for the energy-limited regime,
using the examples defined in Section B 2 above.
4. Niche Overlap
To find out what controls the diversity of the diverse
regime, we varied the niche overlap, which quantifies
the similarity among consumer preferences within the re-
gional species pool. We did this by holding µc fixed, and
varying c1 from its original value of 1 down to a mini-
mum value of 0.12. For each value of c1, we generated 10
ciα matrices, which each defined a regional pool of 200
species. We then repeated the procedure of Section B 2
above for each of these regional pools: initializing 10 wells
with 100 species and running them to the steady state
with the same sequence of propagation and passage steps.
The final richness of each community is plotted in Fig-
ure 4 of the main text as a translucent point, such that
more common richness values are darker. We included
all three examples defined at the end of Section B 2 in
the plot, and colored both examples from the resource-
limited regime blue, while the diverse regime was colored
red.
To examine the beta diversity patterns in each regime,
we initialized 200 wells with 100 randomly chosen species
from the regional pool of 200 species, and propagated
them to steady state following Section B 2 under the three
different choices of w0κ0 and l listed at the end of that
section. To visualize the variation among these commu-
nities, we used the Python package scikit-learn [43] to
compute the first two principle components of the set of
composition vectors in each regime. We then projected
the compositions onto the plane spanned by these vec-
tors, and generated a scatter plot of the results. We also
computed the percentage of the total variance accounted
for by each of these two principal components, and indi-
cated the value in parentheses on each axis.
6. Data Format
The output of all the simulations was saved to a set
of Microsoft Excel spreadsheets, which can be easily im-
ported into Python for analysis using the Pandas pack-
age. Each simulation generated four files: final consumer
populations ('Consumers'), final resource concentrations
('Resources'), a metadata summary ('Parameters'),
and initial conditions ('Initial_State'). The ciα and
Dαβ matrices as well as the mi and wα were pickled into
a binary file ('Realization'). The file names also include
the date on which the data was generated, and a task ID
when multiple files were generated on the same day.
The first column in the consumer and resource tables
is the index of the simulation run. The second and
third columns of the consumer file are the family ID
and species ID, respectively. In the resource file, these
columns contain the class ID and resource ID. The re-
maining columns contain the populations/concentrations
for each well. The consumer populations are in units of
106 cells.
All the parameters that change between runs are in-
cluded in the metadata file ('Parameters'). The first col-
umn of this file is the simulation run index, corresponding
to the index in the consumer and resource files.
The initial conditions file contains the initial popula-
tion sizes for each of the wells.
Appendix C: Robustness of qualitative results
In this Appendix, we test the robustness of our quali-
tative results by modifying the modeling assumptions in
five ways. We have given each way a descriptive name,
which can be used to look up the raw data files from the
supplemental data folder using the file_list.csv table:
• main_dataset is the data from the main text
• type_II uses a Type II functional response, with
K = 20.
• dense_metabolism has a dense metabolic matrix
with d0 = 0.001.
• randomness adds (quenched) random variation to
wα and lα, with standard deviations 0.1 and 0.03,
respectively.
• Gaussian_sampling samples the ciα's from Gaus-
sian distributions, with the same mean 0.11 and
standard deviation 0.3 as the binary matrix used
in the main text.
• Gamma_sampling samples the ciα's from Gamma
distributions, with the same mean and variance.
The following sections describe each of these choices
in more detail. Figures 8, 11, 12 and 13 show the key
plots from the main text along with the new versions
generated under all these modified assumptions. Figures
9 and 10 display another diversity measure not discussed
in the main text: the Simpson Diversity (S. D.). This is
defined analogously to the "effective number of resources
consumed" presented in Equation (6) of the main text:
(cid:34)(cid:88)
(cid:18) Ni
(cid:19)2(cid:35)−1
(C1)
S.D. =
N
i
(cid:80)
where N ≡
i Ni. As discussed in the main text in con-
nection with resource fluxes, this quantity approaches 1
when there is one large Ni ≈ N and all the other popula-
tions are very small. It approaches the number of species
(i.e., the richness) as the biomass distribution becomes
more uniform.
1. Type-II Growth
We chose the Monod parameter K = 20 in the Type-II
growth simulations in order to ensure that at least one
species would survive in the steady state in all simula-
tions. The maximum possible incoming energy flux in
the Type-II model is equal to 0.1K when wα = 1 and
l = 0.9, and this must exceed mi ≈ 1 for a species to
survive. K = 20 provides a maximum flux of 2 in this
case.
2. Metabolic Matrices
The metabolic matrices Dβα are plotted in Figure 3 of
the main text for main_dataset and dense_metabolism
(all other simulations use the same metabolic parame-
ters as main_dataset). We see that d0 = 0.2 leads to
a very sparse matrix, with only a few secreted byprod-
ucts per input resource, while the secretion fractions for
d0 = 0.001 are much more uniform.
16
3. Randomness in wα and lα
To relax the assumption of all the wα's and lα's being
equal, we sampled these two vectors from Gaussian dis-
tributions. We chose the standard deviations of the dis-
tributions to be small enough that both quantities would
almost always be positive, and lα would remain less than
1.
4. Gaussian and Gamma Sampling
Sampling consumer preferences from the continuous
Gaussian and Gamma distributions makes the differen-
tial equations much more stiff than in the binary case. To
ensure stable operation of the integrator, we "passaged"
the cells every 0.1 time units. Each call of the "passage"
method zeros out small negative values of resource con-
centration or consumer population that arise because of
numerical error, in addition to setting small consumer
populations to zero. This high frequency of passaging
made the simulation more computationally intensive, so
we only propagated these simulations for 200 time units.
We computed the root-mean-square difference between
the per-capita growth rates (1/Ni)(dNi/dt) and zero to
check whether the simulations had converged. We found
that all of them had acceptably converged, except for
some of the runs at w0κ0 < 100 in the Gaussian case.
The Gaussian model is unphysical, because almost half of
the consumer preferences are less than 0 for these simula-
tions, and so we decided not to spend more computation
time in pursuit of convergence.
5. Quantification of Nestedness
Almeida-Neto et al. have introduced a quantitative
measure of nestedness, called the "Nestedness metric
based on Overlap and Decreasing Fill," or NODF [2].
Figure 14 shows how the NODF depends on the relative
abundance threshold for the Tara Oceans data, as com-
pared with two null models taken from the Earth Mi-
crobiome Project analysis [49]. Null Model 1 keeps the
richness of each sample the same while randomly alter-
ing the identities of the surviving species. It tells us how
much nestedness we should expect by chance from a set
of samples with the given levels of diversity, in the ab-
sence of any ecological mechanisms. Null Model 2 keeps
the prevalence of each species the same while randomly
assigning it to different samples. This procedure gener-
ates another well-defined family of random matrices with
similar bulk statistics to the original data, but lacks the
operational interpretation of Null Model 1. For this rea-
son, the comparison with Null Model 1 is more meaning-
ful, but we include Null Model 2 for completeness.
For each of the null models and each value of the rela-
tive abundance threshold, we generated 100 random per-
mutations of the data matrix and computed the mean
and standard deviation of their NODF scores. We found
that the actual nestedness exceeds that of Null Model 1
by at least 20 standard deviations for all values of the
threshold. For the relative abundance threshold of 0.5%
employed in Figure 7, the true NODF also exceeds Null
Model 2 by 5.8 standard deviations.
The figure also shows histograms generated with the
two null models from the simulation data of Figure 6A.
The actual NODF (=0.46) is more than 100 standard
deviations above the mean nestedness for both models.
To compute the NODF, we employed the following al-
gorithm, which is implemented in the Community Sim-
ulator package (in the analysis module). Let n be the
number of columns, and m the number of rows in a ma-
trix A. Let Dc be an n × n matrix, such that (Dc)ij = 1
if the sum of column i is greater than the sum of column
j, and 0 otherwise. Similarly, let Dr be an m × m ma-
trix such that (Dr)ij = 1 if the sum of row j is greater
than the sum of row i, and zero otherwise. Let B be the
row-normalized matrix, where each row of A has been
divided by the sum over the row. And let C be the
column-normalized matrix, where each column has been
normalized by the sum over the column. Then the NODF
of the matrix A is
Tr(AT DrB) + Tr(ADcC T )
NODF = 2
n(n − 1) + m(m − 1)
where Tr represents the trace operation.
(C2)
Appendix D: Phase Transition
A phase transition in physics is characterized by a dis-
continuous change in the value of an observable or its
in the
derivative as an intensive parameter is varied,
"thermodynamic limit" of infinite system size.
In an
ecological context, the analog to system size is the num-
ber of possible resource types M , or the initial num-
ber of species S. Several recent works have explored
the analytic computations that become tractable in the
M, S → ∞ limit of various models, while γ ≡ M/S
remains constant (when the model is resource-explicit)
[1, 4, 7, 50]. Taking this limit requires several decisions
about how to scale the rest of the parameters. Our sam-
pling scheme for the ciα and Dαβ matrices, described in
Appendix B, follows the canonical strategy for studying
spin glasses, where the random coupling parameters Jij
are chosen such that the mean and variance are both
proportional to 1/M [40]. The maintenance costs mi, on
the other hand, are sampled from the same distribution
regardless of the value of M . Finally, we note that the
total amount of energy w0κ0 supplied to the system is
an "extensive" parameter, and that the scaling analysis
should be performed with the corresponding "intensive"
parameter w0κ0/M held fixed.
Figure 15 shows how the consumer richness scales with
M for each of the three examples discussed in the main
17
text. The first two examples come from the resource-
limited regime. The "syntrophy-limited" example has
w0κ0/M = 10, l = 0.1, while the "energy-limited" exam-
ple has w0κ0/M = 0.28, l = 0.6. The third, "similarity-
limited" example comes from the diverse regime, with
w0κ0/M = 10, l = 0.9. The richness appears to scale
like M α with exponent α < 1 for the resource-limited
examples, and α = 1 for the diverse example.
If this
scaling holds asymptotically as M → ∞, then the sys-
tem exhibits a true phase transition, with the normalized
richness (richness/M ) vanishing in the resource-limited
regime, while remaining finite in the diverse regime. The
gray line in the right-hand panel illustrates what this
would look like, with a discontinuity in the derivative of
the normalized richness as a function of l or w0κ0/M .
This evidence is by no means conclusive, since we only
have access to a single decade of M values. To reach
three decades of M values would require solving 40,000
coupled ODE's involving matrices ciα and Dαβ with size
20, 000 × 20, 000. Each matrix would thus have 4 × 108
entries, corresponding to 50 MB of memory for binary
entries or 800 MB of memory for floating-point entries.
This computation is feasible but non-trivial, demanding
significantly more attention to how the matrix multipli-
cations are implemented and how the matrices are passed
around in memory. We are currently working on an up-
date to the Community Simulator package that imple-
ments the core computations in PyTorch, which enables
GPU acceleration of matrix multiplication and may allow
for calculations on this scale.
For completeness, Figure 16 shows how four other
natural observables scale with system size. These are
the Simpson and Shannon diversity of the steady-state
consumer and resource abundances [52]. To compute
these quantities, one first obtains relative abundances
α Rα. In terms of these
fi = Ni/(cid:80)
fractions, the Simpson diversity is
i Ni and fα = Rα/(cid:80)
(cid:32)(cid:88)
DSim =
(cid:33)−1
(D1)
and is related to the Inverse Participation Ratio com-
monly analyzed in spin glass problems, while the Shan-
non diversity is
f 2
i
i
(cid:88)
i
(cid:32)
(cid:33)
DSh = exp
−
fi ln fi
,
(D2)
and is simply the exponential of the Shannon entropy of
the distribution (where the sum is taken only over the
species with nonzero abundance). Both of these quanti-
ties are equal to 1 in the limit where a single type domi-
nates the distribution, and equal the total number of sur-
viving types when all the types have the same abundance.
The Simpson and Shannon diversity of the consumers ap-
pear to saturate at a finite values in the large M limit
of the resource-limited regime. When measured in these
ways, the diversity of this regime thus appears to be in-
sensitive to the size of the regional species pool and to
the number of possible resource types, and is controlled
by the energy supply, leakage fraction, and probably also
the sparsity of the Dαβ matrix.
18
19
FIG. 8. Richness vs. w0κ0 and (cid:104)lα(cid:105) We generated 200 species, initialized 10 communities of 100 species each from this pool,
and ran the dynamics to steady state under different combinations of w0κ0 and (cid:104)lα(cid:105), for each of the six model choices listed at
the end of Section B 2. The color of each square indicates the mean number of non-extinct species at the end of the simulation,
over all 10 communities at each combination of w0κ0 and (cid:104)lα(cid:105).
FIG. 9. Simpson Diversity vs. w0κ0 and (cid:104)lα(cid:105). Simpson Diversity was computed according to Equation (C1), using the
same data as Figure 8.
Leakage Fraction l101001000Supplied Energy Flux w00main_datasetLeakage Fraction lSupplied Energy Flux w00type_IILeakage Fraction lSupplied Energy Flux w00dense_metabolism0.00.10.20.30.40.50.60.70.80.9Leakage Fraction l101001000Supplied Energy Flux w00randomness0.00.10.20.30.40.50.60.70.80.9Leakage Fraction lSupplied Energy Flux w00Gaussian_sampling0.00.10.20.30.40.50.60.70.80.9Leakage Fraction lSupplied Energy Flux w00Gamma_sampling051015202530051015202530051015202530051015202530051015202530051015202530Leakage Fraction l101001000Supplied Energy Flux w00main_datasetLeakage Fraction lSupplied Energy Flux w00type_IILeakage Fraction lSupplied Energy Flux w00dense_metabolism0.00.10.20.30.40.50.60.70.80.9Leakage Fraction l101001000Supplied Energy Flux w00randomness0.00.10.20.30.40.50.60.70.80.9Leakage Fraction lSupplied Energy Flux w00Gaussian_sampling0.00.10.20.30.40.50.60.70.80.9Leakage Fraction lSupplied Energy Flux w00Gamma_sampling04812162004812162004812162004812162004812162004812162020
FIG. 10. Richness (blue solid) and Simpson Diversity (red dotted) for cuts through the heat map. All six
modeling choices are plotted, going from dark to light in the order listed in the text. Error bars are plus and minus one
standard deviation, where the standard deviation is computed over the 10 parallel communities at each set of parameter values.
02004006008001000Supplied Energy Flux w00051015202530High Leakage (l=0.9)0.00.20.40.60.8Leakage Fraction lHigh Energy Supply (w00=1000)21
FIG. 11. Rank-abundance curves for three representative examples. The abundance is plotted in absolute number of
cells, as opposed to the relative abundance of the main text, and includes all species, with no truncation. The lower limit of
the vertical axis is set to 1 cell, which is the smallest possible population size once the Passage method described in Appendix
B has generated integer population values. The three examples are the same ones highlighted in Figure 2 of the main text, and
listed in Section B 2 above.
Rank100102104106108Absolute Abundancemain_datasetSyntrophy-LimitedRankEnergy-LimitedRankSimilarity-LimitedRank100102104106108Absolute Abundancetype_IIRankRankRank100102104106108Absolute Abundancedense_metabolismRankRankRank100102104106108Absolute AbundancerandomnessRankRankRank100102104106108Absolute AbundanceGaussian_samplingRankRank0102030Rank100102104106108Absolute AbundanceGamma_sampling0102030Rank0102030Rank22
FIG. 12. Effective number of resources consumed. The effective number of resources consumed M eff
is computed as
described in the main text for the same three examples. The y axis indicates the total number of species falling into each bin
from the combination of all 10 parallel communities in each example.
i
FIG. 13. Resource susceptibility. Histograms of the logarithm log10 ∂ ¯Rα/∂κα are plotted for the same three examples.
The y axis indicates the total number of species falling into each bin from the combination of all resources for all 10 parallel
communities in each example, excepting the externally supplied resource α = 0.
020406080Number of Speciesmain_datasetEnergy-LimitedSyntrophy-LimitedSimilarity-Limitedtype_IIdense_metabolism051015020406080Number of Speciesrandomness051015Effective Number of Resources ConsumedGaussian_sampling051015Gamma_samplingFrequencymain_datasetEnergy-LimitedSyntrophy-LimitedSimilarity-Limitedtype_IIdense_metabolism3210Frequencyrandomness3210Log Resource SusceptibilityGaussian_sampling3210Gamma_sampling23
FIG. 14. Quantification of nestedness. (a) Nestedness as quantified by NODF [2] of the Tara Oceans data, for different
values of the relative abundance threshold. See Appendix C 5 for the full quantification algorithm. Also shown are the mean
NODF over 100 samples from two null models, with error bars representing ±1 standard deviation. Null Model 1 keeps the
richness of each sample the same while randomly altering the identities of the surviving species. Null Model 2 keeps the
prevalence of each species the same while randomly assigning it to different samples. (b) Histograms of 1000 samples from
each null model at the relative abundance threshold of 0.005 used in the main text, and indicated with a black line in the first
panel. (c), (d) Null models for comparison with the NODF value of 0.46 obtained for the simulation data. The two panels
show histograms of 1,000 samples from null models 1 and 2, respectively.
FIG. 15. Scaling of consumer richness with system size. Left: Community richness in the three examples of Section
B 2 as a function of the number of resource types M (with the rest of the parameters scaled as described in Appendix B). The
thin black line shows the slope corresponding to a linear scaling with M . Error bars are standard deviations over 100 samples
per point. Right: Normalized richness as a function of leakage fraction at high energy supply w0κ0/M = 10. The top red line
comes from the simulations of Figure 2 of the main text (M = 100), and the middle blue line comes from new simulations with
M = 500. The bottom gray line illustrates a possible M → ∞ limit, which would correspond to a continuous phase transition.
Error bars are standard deviations over 10 samples per point.
(a)(b)(c)(d)102103M100101102RichnessSimilarity-LimitedEnergy-LimitedSyntrophy-Limited0.00.10.20.30.40.50.60.70.80.9Leakage0.000.050.100.150.200.250.300.35Richness/MM = ∞?M = 100M = 50024
FIG. 16. Scaling of other observables. Diversity of steady-state consumer and resource abundances in the three examples
of Section B 2, as a function of the number of resource types M (with the rest of the parameters scaled as described in Appendix
B). The thin black line shows the slope corresponding to a linear scaling with M . Error bars are standard deviations over 100
samples per point.
FIG. 17. Flux network topology. We generated flux networks like those of Figure 3 for both the sparse ('main dataset') and
the dense ('dense metabolism') metabolic matrices shown there. We pruned the graphs as in the main text figure by removing
edges with flux less than 1% of maximum flux in the network. For each condition, we determined how many of the ten replicates
had acyclic graphs, which could thus be topologically ordered like the left-hand panels of the figure. The color of each square
represents the fraction of acyclic graphs for the given values of l and w0κ0.
100101102Consumer DiversitySyntrophy-LimitedRichnessSimpsonShannonEnergy-LimitedSimilarity-Limited102103M100101102Resource Diversity102103M102103M00.20.40.60.8Leakage Fraction l101001000Supplied Energy Flux w00main_dataset00.20.40.60.8Leakage Fraction ldense_metabolism0.00.20.40.60.81.0Fraction of wells with no metabolic cycles0.00.20.40.60.81.0Fraction of wells with no metabolic cycles25
FIG. 18. Number of available resources. The mean number of resources (over all ten initial communities) with abundances
Rα above 0.1 is plotted for each of the conditions summarized in Figure 2. This threshold is the level at which an average
consumer equally consuming 10 resources would be able to satisfy its full maintenance cost.
00.20.40.60.8Leakage Fraction l101001000Supplied Energy Flux w0020406080Mean Number of Available Resources26
FIG. 19. Maintenance costs mi. The maintenance cost mi for each species is plotted against the relative abundance of that
species in its community, for each of the simulations in Figure 2. The ten panels contain data for each of the ten initial pools
of species. In each panel, each species has 100 data points, one for each of the 100 combinations of l and w0κ0. The points
are colored by the richness of the steady-state community, with blue being least diverse and red the most diverse. Every point
in a given panel corresponds to a species that was initially present in the community, so points on the mi axis where all the
relative abundances vanish are species that never survive to steady state in any of the sampled conditions.
|
1705.02624 | 1 | 1705 | 2017-05-07T13:51:39 | State assignment problem in systems biology and medicine: on the importance of state interaction network topology | [
"physics.bio-ph",
"cond-mat.dis-nn",
"cond-mat.stat-mech",
"physics.med-ph"
] | A fundamental problem in medicine and biology is to assign states, e.g. healthy or diseased, to cells, organs or individuals. State assignment or making a diagnosis is often a nontrivial and challenging process and, with the advent of omics technologies, the diagnostic challenge is becoming more and more serious. The challenge lies not only in the increasing number of measured properties and dynamics of the system (e.g. cell or human body) but also in the co-evolution of multiple states and overlapping properties, and degeneracy of states. We develop, from first principles, a generic rational framework for state assignment in cell biology and medicine, and demonstrate its applicability with a few simple theoretical case studies from medical diagnostics. We show how disease-related statistical information can be used to build a comprehensive model that includes the relevant dependencies between clinical and laboratory findings (signs) and diseases. In particular, we include disease-disease and sign-sign interactions. We then study how one can infer the probability of a disease in a patient with given signs. We perform comparative analysis with simple benchmark models to check the performances of our models. This first principles approach, as we show, enables the construction of consensus diagnostic flow charts and facilitates the early diagnosis of disease. Additionally, we envision that our approach will find applications in systems biology, and in particular, in characterizing the phenome via the metabolome, the proteome, the transcriptome, and the genome. | physics.bio-ph | physics |
State assignment problem in systems biology and medicine: on
the importance of state interaction network topology
Abolfazl Ramezanpoura,b, Alireza Mashaghia,c,∗
aLeiden Academic Centre for Drug Research,
Faculty of Mathematics and Natural Sciences,
Leiden University, Leiden, The Netherlands
bDepartment of Physics, University of Neyshabur, Neyshabur, Iran
cHarvard Medical School, Harvard University, Boston, Massachusetts, USA and
∗Correspondence to [email protected]
(Dated: July 12, 2021)
Abstract
A fundamental problem in medicine and biology is to assign states, e.g. healthy or diseased, to
cells, organs or individuals. State assignment or making a diagnosis is often a nontrivial and chal-
lenging process and, with the advent of omics technologies, the diagnostic challenge is becoming
more and more serious. The challenge lies not only in the increasing number of measured proper-
ties and dynamics of the system (e.g. cell or human body) but also in the co-evolution of multiple
states and overlapping properties, and degeneracy of states. We develop, from first principles, a
generic rational framework for state assignment in cell biology and medicine, and demonstrate its
applicability with a few simple theoretical case studies from medical diagnostics. We show how
disease-related statistical information can be used to build a comprehensive model that includes
the relevant dependencies between clinical and laboratory findings (signs) and diseases. In par-
ticular, we include disease-disease and sign-sign interactions. We then study how one can infer
the probability of a disease in a patient with given signs. We perform comparative analysis with
simple benchmark models to check the performances of our models. This first principles approach,
as we show, enables the construction of consensus diagnostic flow charts and facilitates the early
diagnosis of disease. Additionally, we envision that our approach will find applications in systems
biology, and in particular, in characterizing the phenome via the metabolome, the proteome, the
transcriptome, and the genome.
1
I.
INTRODUCTION
Human body as a whole or in part may adopt various states, like a Rubik's Cube. Home-
ostatic mechanisms, medical interventions and aging all involve evolution from certain body
states to others. Similarly, evolution of states is commonly seen in cells that constitute
our bodies. Immune cells for example can manifest substantial plasticity and develop into
distinct phenotypes with different functions [2, 3]. Identifying dysfunctional states in cells
is in many ways similar to identifying diseases in organisms and is confronted by similar
difficulties. State assignment, as we describe below, is often a nontrivial and challenging
process and in many cases it is hard to diagnose the state of a cell or conditions of a pa-
tient. The progress in systems biology and availability of large data has made diagnostics
even more challenging. For instance, mass cytometric analysis of immune cells has led to
identification of many new cell subtypes (states) [4]. Metabolomic analysis of cells and body
fluids revealed a large number of biomarkers and disease subtypes [5]. There is a huge need
in the fields of cell biology, immunology, clinical sciences and pharmaceutical sciences for
approaches to identify states, assigning states and characterizing co-emerging or co-existing
states. Moreover, it is often important to be able to identify emerging states even before
they are fully evolved. From physics point of view, it is interesting to yield a generic un-
derstanding of the state assignment problem in cell biology or medicine (although specific
details might be crucial in each context in practice). Without loss of generality, in what fol-
lows we focus on medical diagnostics and draw a simple picture that captures many generic
aspects of assignment problems in medicine and systems cell biology.
Decision-making is at the heart of medicine. Decisions are made at various stages in
clinical practice, particularly during diagnostic investigations and when assigning the find-
ings to a disease [6, 7]. Diagnostic strategies are typically available in the form of clinical
algorithms and flow charts that define the sequence of actions to be taken to reach a diag-
nosis. The diagnosis itself is typically made based on consensus diagnostic criteria [8, 9]. In
addition, there are a number of clinical decision support systems and software systems that
are used to assign findings (symptoms and signs) to disease conditions. The most commonly
used technologies are WebMD Symptom Checker, Isabel Symptom Checker, DXplain and
Internist[10]. These algorithms compute the most likely disease that is associated with a
given set of findings by using only a small part of the existing probabilistic data on findings
2
and diseases. Internist, which is one of the most sophisticated systems, relies on two pa-
rameters, the probability of a finding given a disease and the probability of a disease given
a finding [11]. These technologies inform us if a patient satisfies the criteria of a disease but
do not provide guidance on how to approach a patient and mostly ignore the interactions
between diseases.
Currently, we lack a solid conceptual framework for medical diagnostics. As a conse-
quence, there is no consensus on the diagnostic flow charts available today, and clinicians
differ widely in their approaches to patients. Here, we take a step towards solving this prob-
lem by formulating first principles medical diagnostics. We evaluate the performance of the
platform and discuss how one can optimize it. Using simple theoretical examples, we show
how including relevant statistical data and often-ignored inherent disease-disease linkages
significantly reduces diagnostic errors and mismanagement and enables the early diagnosis
of disease.
The problem of associating a subset of observed signs (clinical signs/symptoms and labo-
ratory data) with specific diseases was easy if we could assume the findings originate from a
single disease, we had clear demonstrations for the diseases, and inter-sign and inter-disease
interactions were negligible. In practice, however, we typically have no certain relationships
that connect signs to diseases, and one often must address interference effects of multiple
diseases; in the early stages of a disease, we do not even have sufficient findings to make a
definite decision [12 -- 14]. There are a number of studies that have attempted to quantify
such dependencies under uncertainty and obtain estimations for the likelihood of diseases
given a subset of findings [11, 15 -- 20]. An essential simplifying assumption in these studies
was that only one disease is behind the findings (exclusive diseases assumption). Among
recent developments, we should mention Bayesian belief networks, which provide a proba-
bilistic framework to study sign-disease dependencies [21 -- 24]. These models are represented
by tables of conditional probabilities that show how the state of a node (sign or disease)
variable in an acyclic directed graph depends on the state of the parent variables. Here,
it is usually assumed that the signs are conditionally independent of one another given a
disease hypothesis and that diseases are independent of one another after marginalizing
over the sign variables (marginally independent diseases)[22]. In other words, there exist
no causal dependencies or interactions (directed links in the acyclic graph) that connect
the signs/diseases. In this study, however, we shall pursue a more principled approach to
3
highlight the significance of direct disease-disease and sign-sign interactions (dependencies).
Evidences are rapidly growing to support the existence of such interactions [25 -- 32]. Our
approach is of course computationally more expensive than the previous approaches, but it
shows how different types of interactions could be helpful in the course of diagnosis. Addi-
tionally, because of recent developments in related fields [33 -- 36], we now have the necessary
concepts and tools to address difficulties in more sophisticated (realistic) problems of this
type. This study does not involve usage of real medical data, which is by the way fun-
damentally incomplete at this moment for such modeling; however, it provides a rationale
as to why certain often-neglected statistical information and medical data can be useful in
diagnosis and demonstrates that investments in collecting such data will likely pay off.
II. PROBLEM STATEMENT
Consider a set of ND binary variables D = {Da = 0, 1 : a = 1, . . . , ND}, where Da = 0, 1
shows the absence or presence of disease a. We have another set of NS binary variables
S = {Si = ±1 : i = 1, . . . , NS} to show the values of sign (symptom) variables.
Suppose we have the conditional probability of symptoms given a disease hypothesis,
P (SD), and prior probability of diseases P0(D). Then, the joint probability distribution
of sign and disease variables reads as P (S; D) ≡ P (SD)P0(D). We shall assume, for
simplicity, that the probability distributions describe the stationary state of the variables.
The distributions may be subject to environmental and evolutionary changes and may also
change in the course of the disease. Here, we limit our study to the time scales that are
smaller than the dynamical time scale of the model and leave the temporal dynamics for a
future study. In addition, we assume that we are given sufficient statistical data, e.g., the true
marginal probabilities Ptrue(Si, SjD), to reconstruct simple models of the true probability
III A, we propose
distribution [37]. This is indeed the first part of our study:
In Sec.
statistical models of sign and disease variables, and employ efficient learning algorithms to
compute the model parameters, given the appropriate data. Fortunately, recent advances in
machine learning and inference techniques enable us to work with models that involve very
large number of variables [36, 38 -- 45].
Let us assume that a subset O = {i1, i2, . . . , iNO} of the sign variables has been observed
with values So, and size NO = O. We will use U for the remaining subset of unobserved
4
signs with values which are denoted by Su. Then, the likelihood of disease variables given
the observed signs is:
The most likely diseases are obtained by maximizing the above likelihood:
L(DSo) ≡XSu
P (S; D).
DM L = arg max
D
log L(DSo).
(1)
(2)
Here, we are interested in the posterior probability marginals P (Da = 0, 1) of the disease
variables. The marginal probability P (Sj = ±1) of an unobserved sign, and the most likely
signs, are obtained from the following distribution:
M(SuSo) ≡XD
P (S; D),
SM L = arg max
Su
log M(SuSo).
(3)
The main task in the second part of our study is computation of the sign and disease marginal
probabilities
P (Da) ∝ X{Db:b6=a}
L(DSo),
P (Sj) ∝ X{Sk:k∈U \j}
M(SuSo)
j ∈ U.
(4)
In general, computing the exact values of these marginals is a hard problem. However, one
can find highly accurate approximation methods developed in the artificial intelligence and
statistical physics communities to address such computationally difficult problems [33, 34,
46 -- 49]. In Sec. III B, we propose an approximate message-passing algorithm for inferring
the above information in a large-scale problem.
Finally, the last and main part of our study is devoted to the problem of choosing a finite
sequence of unobserved signs for observation, which maximizes an appropriate objective
functional of the sequence of observations. In principle, the objective function should be
designed to approach the right diagnosis in a small number of observations. To this end, we
assign larger values to the objective function if the observations result to larger polarization
in the disease probabilities; obviously, it is easier to decide if disease a is present or not when
the marginal probability P (Da) is closer to 0 or 1 (more polarized). Computing such an
objective functional of the disease probabilities for a given sequence of observations is not an
easy task. We have to consider also the stochastic nature of the observations; we know the
sign probabilities P (Sj), but, we do not know a priori the value Sj of an unobserved sign,
5
which is chosen for observation. To take into account this uncertainty, we shall work with
an objective function which is averaged over the possible outcomes of the observation. More
precisely, the above diagnosis problem is a multistage stochastic optimization problem, a
subject that has been extensively studied in the optimization community [35, 50 -- 52].
Suppose we are to observe T ≤ NS − NO signs with an specified order OT ≡ {j1, . . . , jT};
there are (NS − NO)!/(T !(NS − NO − T )!) different ways of choosing T signs from NS −
NO ones, and T ! different orderings of the selected signs to identify such a sequence of
observations. Therefore, the number of possible sequences grows exponentially with T .
Add to this the computational complexity of working with an objective functional of the
sequence of observations, which has to be also averaged over the stochastic outcomes of the
observations. In Sec. III C, we present simple heuristic and greedy algorithms to address the
above problem, and leave a detailed study of the multistage stochastic optimization problem
for future.
III. RESULTS
A complete description of a collection of stochastic variables, like the sign and disease
variables, is provided by the joint probability distribution of the variables P (S; D). Having
the probability distribution (model) that describes a system of interacting variables does not,
however, mean that one can readily extract useful statistical information from the model. In
fact, both the model construction and the task of extracting information from the model are
computationally hard, with computation times that in the worst cases grow exponentially
with the number of involved variables [53 -- 57]. In the following, we address the above sub-
problems in addition to the main problem of optimizing an appropriate objective functional
of observations, which are made during the course of diagnosis.
A. Learning the model: Maximum entropy principle
Let us assume that we are given the marginal probabilities Ptrue(Si, SjD) of the true
conditional probability Ptrue(SD). Then, we use the maximum entropy principle to con-
struct an appropriate model P (S; D) ≡ P (SD)P0(D) of the sign and disease variables
[58 -- 60]. Here, P0(D) is the prior probability of the diseases depending on the age, gender,
6
P(SD)
P (D)
0
b
a
P(S0)
j
i
α
Prior Disease Interaction Sign Leak
FIG. 1. The interaction graph of disease variables (left circles) and sign variables (right circles)
related by Ma one-disease and Mab two-disease interaction factors (middle squares) in addition to
interactions induced by the leak probability (right square) and the prior probability of diseases
(left square). In general, an interaction factor α = a, ab is connected to kα signs and lα diseases.
and other characteristics.
In the following, we simply take a product prior distribution,
P0(D) = Qa P0(Da). In the absence of any prior information, the above probability dis-
tribution is uniform. The conditional probability P (SD) represents all the sign/disease
interactions that are allowed by the maximum entropy principle,
P (SD) =
1
Z(D)
exp Xi
hi(D)Si +Xi<j
Jij(D)SiSj! .
(5)
Here Z(D) is the normalization (or partition) function.
In practice, we are given only a small subset of the conditional probabilities, for instance,
Ptrue(Si, SjonlyDa) and Ptrue(Si, SjonlyDa, Db). The former is the probability that signs i
and j take values (Si = ±1, Sj = ±1) conditioned to the presence of disease a and the absence
of all other diseases. The latter conditional probabilities are defined similarly. Therefore,
we have to consider only interactions between a small number of disease variables. To this
end, we expand the model parameters,
hi(D) = h0
ha
i +Xa
ij +Xa
i Da +Xa<b
ijDa +Xa<b
J a
Jij(D) = J 0
hab
i DaDb + · · · ,
J ab
ij DaDb + · · · ,
(6)
(7)
and keep only the leading terms of the expansion. Putting all together, given the above
7
information, we rewrite
P (SD) =
1
Z(D)
φ0(S) ×Ya
φa(SDa) ×Ya<b
φab(SDa, Db).
(8)
Here, φ0 is responsible for the leak probabilities P (Snodisease), to account for the missing
disease information and other sources of error [22, 24]. In the following, we assume that
local sign fields are sufficient to produce an accurate representation of the leak probabilities,
i Si) . The other interaction factors, φa and φab, are present only if the
associated diseases are present; they are written in terms of local sign fields and two-sign
i.e., φ0 = exp(Pi K 0
interactions:
K a
φa = exp(Da[Xi
φab = exp(DaDb[Xi
K a
i Si +Xi<j
i Si +Xi<j
K ab
ijSiSj]),
K ab
ij SiSj]).
(9)
(10)
Figure 1 shows a graphical representation of the model, which has Ma one-disease and
Mab two-disease interaction factors, each of which is connected to ka and kab sign variables,
respectively. From the above model, we obtain the simpler one-disease-one-sign (D1S1)
model in which we have only the one-disease factors (i.e., Mab = 0) and local sign fields (i.e.,
K a
ij = 0). In a two-disease-one-sign model (D2S1), we have both the one- and two-disease
factors, but only the local sign fields.
In the same way, we define the one-disease-two-
sign (D1S2) and two-disease-two-sign (D2S2) models. In the following, unless otherwise
mentioned, we shall work with the fully connected graphs with parameters: Ma = ND, ka =
NS for the D1S1 and D1S2 models, and Ma = ND, Mab = ND(ND − 1)/2, ka = kab = NS
for the D2S1 and D2S2 models. Moreover, the interaction factors in the D1S2 and D2S2
models include all the possible two-sign interactions in addition to the local sign fields.
To obtain the model parameters (K 0
), we start from the conditional
marginals Ptrue(Sinodisease). This information is sufficient to determine the couplings K 0
from the following consistency equations:
, and K a,ab
i , K a,ab
ij
i
i
Ptrue(Sinodisease) = X{Sj :j6=i}
P (SD = 0)
∀i.
(11)
If we have Ptrue(Si, SjonlyDa), then in principle we can find K a
consistency equations, assuming that we already know the K 0
i and K a
ij from similar
i . Note that P (Si, SjonlyDa)
8
is different from P (Si, SjDa), which is conditioned only on the value of disease a. In the
same way, having the Ptrue(Si, SjonlyDa, Db) allow us to find the couplings K ab
i and K ab
ij ,
In general, the problem of finding the couplings from the above conditional
and so on.
probabilities is computationally expensive. However, there are many efficient approximate
methods that enable us to find good estimations for the above couplings given the above
conditional probabilities [36, 38 -- 41, 43, 45]. The reader can find more details about the
models in Appendix A, where we provide a very simple learning algorithm, which is based
on the Bethe approximation, for estimating the model parameters, given the above marginal
probabilities.
B. Computing the marginal probabilities: An approximate inference algorithm
Let us consider a simple benchmark model to check the performances of the above con-
structed models. As the benchmark, we take the true conditional probability
Ptrue(SD) =
1
Ztrue(D)
e−H(S,S∗(D)),
(12)
where S∗(D) gives the symptoms of hypothesis D; we choose these symptoms randomly and
i (D))2/4
is the Hamming distance (number of different signs) of the two sign configurations. Note
uniformly from the space of sign variables. Moreover, H(S, S∗(D)) ≡PNS
i=1(Si − S∗
that there is no sign-sign interaction in the above true model. Therefore, given the true con-
ditional marginals, we can exactly compute the model parameters, as described in Appendix
A.
For small numbers of sign/disease variables, we can use an exhaustive inference al-
gorithm to compute the exact marginal probabilities. Figure 2 displays the sensitivity
P (Si = +1Da = 1) and specificity P (Si = −1Da = 0) of the diseases [24], which have
been obtained by the one-disease-one-sing (D1S1) and two-disease-one-sing (D2S1) models
for a typical realization of the symptoms S∗(D) in the true model. As the figure shows, both
the probabilities are in average closer to the true values in the D2S1 model. This computa-
tion is intended to exhibit the high impact of two-disease interactions on the behavior of the
marginal probabilities. We will soon see that these large effects of interactions can indeed
play a constructive role also in the process of diagnosis.
To infer the marginal probabilities of the models for larger number of sign/disease vari-
9
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.3
0.8
0.7
0.6
0.5
0.4
0.3
0.2
y
t
i
v
i
t
i
s
n
e
s
y
t
i
c
i
f
i
c
e
p
s
0.35
0.4
0.45
0.5
0.55
0.6
0.65
0.1
0.35
0.4
0.45
0.5
0.55
0.6
0.65
0.7
FIG. 2. Sensitivity, P (Si = +1Da = 1), and specificity, P (Si = −1Da = 0), of the D1S1 (open
circles) and D2S1 (filled circles) models versus the true values for a typical realization of the
exponential true model. The model parameters are obtained from the conditional marginals of the
true model. Here we have ND = 5 diseases, NS = 20 signs.
ables, we resort to the Bethe approximation and the Belief-Propagation algorithm [34, 47].
First, we suggest an approximate expression for the normalization function Z(D), which
appears in the denominator of the conditional probability distribution P (SD). In words,
we consider this non-negative function of the diseases to be a probability measure, and we
approximate this measure by a factorized probability distribution, using its one- and two-
variable marginal probabilities (see Appendix B). This approximation enables us to employ
an efficient message-passing algorithm such as belief propagation for computing statistical
properties of the above models. As mentioned before, we shall assume that the prior prob-
ability P0(D) can also be written in an appropriate factorized form. The quality of our ap-
proximations depends very much on the structure of the interaction factors and the strengths
of the associated couplings in the models. The Bethe approximation is exact for interaction
10
graphs that have a tree structure. This approximation is also expected to work very well
in sparsely connected graphs, in which the number of interaction factors (Ma, Mab) and the
number of signs associated with an interaction factor (ka, kab) are small compared with the
total number of sign variables. In Appendix B we display the relative errors in the marginal
signs/diseases probabilities that were obtained by the above approximate algorithm. The
time complexity of our approximate inference algorithm grows linearly with the number
of interaction factors and exponentially with the number of variables that are involved in
such interactions; with ND = 500, NS = 5000, Ma = 500, Mab = 1000, ka = 10, kab = 5, the
algorithm takes approximately one minute of CPU time on a standard PC to compute the
local marginals. We recall that the INTERNIST algorithm works with 534 diseases and
approximately 4040 signs (or manifestations), with 40740 directed links that connect the
diseases to the signs [22].
C. Optimization of the diagnosis process: A stochastic optimization problem
Suppose that we know the results of NO observations (medical tests), and we choose
another unobserved sign j ∈ U for observation. To measure the performance of our decision,
we may compute deviation of the disease probabilities from the neutral values (or "disease
polarization") after the observation:
DP (j) ≡ 1
ND Xa (cid:18)P (Da = 1) −
1
2(cid:19)2!1/2
.
(13)
One can also add other measures such as the cost of observation to the above function.
In a two-stage decision problem, we choose an unobserved sign for observation, with the
aim of maximizing the averaged objective function E(j) ≡ hDP (j)iO. Note that before
doing any real observation, we have access only to the probability of the outcomes P (Sj);
the actual or true value of an unobserved sign becomes clear only after the observation. That
is why here we are taking the average over the possible outcomes, which is denoted by h·iO.
One can repeat the two-stage problem for T times to obtain a sequence of T observations:
each time an optimal sign is chosen for observation followed by a real observation, which
reveals the true value of the observed sign.
In a multistage version of the problem, we want to find an optimal sequence of decisions
OT = {j1,· · · , jT}, which maximizes the following objective functional of the observed
11
signs: E[OT ] ≡ PT
t=1hDP (jt)iO. Here, at each step, the "observed" sign takes a value that
is sampled from the associated marginal probability P (Sj). This probability depends on the
model which we are working with. Note that here we are interested in finding an optimal
sequence of observations at the beginning of the process before doing any real observation. In
other words, in such a multistage problem, we are doing an "extrapolation" or "simulation"
of the observation process without performing any real observation. In practice, however,
we may fix the sequence of observations by a decimation algorithm:
i.e., we repeat the
multistage problem for T times, where each time we observe the first sign suggested by the
output sequence, and reduce the number of to-be-observed sings by one.
In the following, we consider only simple approximations of the multistage problem; first
we reduce the difficult multistage problem to simpler two-stage problems. More precisely,
at each time step, we choose an unobserved sign jt, which results to the largest disease
polarization hDP (jt)iO, for observation (greedy strategy). Then, we consider two cases: (I)
we perform a real observation to reveal the true value of the suggested sign for observation,
(II) we treat the suggested sign variable for observation as a stochastic variable with values
that are sampled from the associated marginal probability.
Let us start from the case in which we observe the true values of the signs chosen for
observation. Once again, we take the exponential benchmark model given by Eq. 12 as the
true model. We use the conditional marginals extracted from this true model to construct
the simple one-disease-one-sign (D1S1) and two-disease-one-sign (D2S1) models. Suppose
that we are given a disease hypothesis D and the associated symptoms S∗(D). We start
from a few randomly chosen observed sings from the set of symptoms. Then, at each time
step t, we compute the inferred sign probabilities P (Sj), and use the greedy strategy to
choose an unobserved sign for observation. The observed sign at each step takes the true
value given by S∗(D). To see how much the disease probabilities obtained from the models
are correlated with the true hypothesis D, after each observation, we compute the following
overlap function (or "disease likelihood"):
DL(t) ≡
1
ND Xa
(2Da − 1)(cid:18)P (Da = 1) −
1
2(cid:19) .
(14)
Besides the magnitude, our decisions also affect the way that the above quantity behaves
with the number of observations.
In Fig. 3 we see how the above overlap function, DL(t), behaves for cases in which
12
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
0.5
0.4
0.3
0.2
0.1
0
0.5
0.4
0.3
0.2
0.1
0
(a)
0
2
(b)
0
2
D1S1 (G)
D2S1 (G)
D1S1 (Random)
D2S1 (Random)
6
4
number of observations
8
10 12 14 16
D1S1 (G)
D2S1 (G)
D1S1 (Random)
D2S1 (Random)
6
4
number of observations
8
10 12 14 16
FIG. 3. Overlap of the inferred disease marginals with the true hypothesis for the exponential
benchmark model. The data are for the cases in which only one (a) or two (b) diseases are present.
The model parameters are obtained from the conditional marginals of the true model. There are
ND = 5 diseases, NS = 20 signs, and the algorithm starts with NO = 3 observed signs for a
randomly selected hypothesis D. An unobserved sign is chosen for observation by the greedy (G)
or random strategy using the inferred probabilities, and the observed sign takes the true value
given by S∗(D). The data are results of averaging over 1000 independent realizations of the true
model and the observation process.
only one or two diseases are present in D (see also Appendix C). For comparison, we also
show the results obtained by a random strategy, where an unobserved sign is chosen for
observation randomly and uniformly from the subset of unobserved signs. The number
of sign/disease variables here is small enough to allow for an exact computation of all the
marginal probabilities. It is seen that both the D1S1 and D2S1 models work very well when
only one disease is present and all the other diseases are absent. The D1S1 model already
fails when two diseases are present in the hypothesis, whereas the other model can still find
13
the right diseases. However, we observe that even the D2S1 model gets confused when
there are more than two diseases in the hypothesis; in such cases, we would need to consider
more sophisticated models with interactions involving more than two diseases. Moreover, we
observe that the difference in the performances of the greedy and random strategies decreases
as the number of involved diseases increases. In Appendix C, we observe similar behaviors
for a more complex benchmark model Ptrue(SD) ∝ 1/(1 + H(S, S∗(D))), including also the
sign-sign interactions.
Next, we consider the case of simulating the diagnosis process without doing any real ob-
servation. Here, we assume that an observed sign takes a value which is sampled from the as-
sociated marginal probability P (Sj) at that time step. For comparison with the greedy strat-
egy, we also introduce two other strategies for choosing an unobserved sign for observation.
A naive strategy is to choose the most positive (MP) sign jmax = arg maxj{P (Sj = +1)} for
observation (MP strategy); jmax is the sign with the maximum probability of being positive.
In the early stages of the diagnosis, this probability is probably close to zero for most of the
signs. So, it makes sense to choose the most positive sign for observation to obtain more
information about the diseases. A more complicated strategy works by first computing the
conditional probabilities P (SjDM L) for the maximum likelihood hypothesis DM L, and then
selecting the most positive sign for observation (MPD strategy).
To have a more general comparison of the constructed models, in the following, we assume
, and K a,ab
ij
i
that the model parameters (K a,ab
) are iid random numbers uniformly distributed
in an appropriate interval of real numbers. The leaky couplings are set to K 0
i = −1, which
correspond to small sign probabilities P (Si = 1nodisease) ≃ 0.05. We assume that all the
possible one-disease and two-disease interaction factors are present in the models. Moreover,
inside each factor we have all the possible two-sign interactions in addition to the local sign
fields. As before, the prior disease probabilities P0(Da) are uniform probability distributions.
Figure 4 shows the performances for different models and strategies with a small number
of sign/disease variables. Here, the "disease likelihood" gives the overlap of the disease
probabilities with the maximum likelihood hypothesis DM L of the models. Moreover, all
the quantities are computed exactly. We see that in this case the average performance of
the greedy strategy is close to that of the MPD strategy at the beginning of the process.
For larger number of observations, the greedy performance degrades and approaches that of
the MP strategy.
14
n
o
i
t
l
a
z
i
r
a
o
p
e
s
a
e
s
d
i
n
o
i
t
a
z
i
r
a
o
p
l
e
s
a
e
s
d
i
0.5
0.45
0.4
0.35
0.3
0.25
0.5
0.45
0.4
0.35
0.3
D1S1
D2S1
D1S2
D2S2
(a)
0
2
4
6
8
10 12 14 16
number of observations
(c)
D2S2: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
0.5
0.45
0.4
0.35
0.3
0.25
0.5
0.45
0.4
0.35
0.3
0.25
D1S1
D2S1
D1S2
D2S2
(b)
0
2
4
6
8
10 12 14 16
number of observations
(d)
D2S2: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
FIG. 4. Diagnostic performance of the models vs the number of observations for a small number of
sign/disease variables. (top) The exact disease-polarization (a) and disease-likelihood (b) obtained
by the MP strategy in the one-disease-one-sign (D1S1), two-disease-one-sign (D2S1), one-disease-
two-sign (D1S2), and two-disease-two-sign (D2S2) models. There are ND = 5 diseases, NS = 20
signs, and the algorithm starts with NO = 4 observed signs with positive values. The couplings
in the interaction factors are iid random numbers distributed uniformly in the specified intervals:
i = −1, K a,ab
K 0
(c) and likelihood (d) of the diseases obtained by the MP, greedy (G), MPD, and random strategies,
ij ∈ (−1, +1)/√NS. (bottom) Comparing the exact polarization
∈ (−1, +1), K a,ab
i
for the D2S2 model. The data are results of averaging over 1000 independent realizations of the
model parameters.
The models with disease-disease and sign-sign interactions exhibit larger polarizations of
the disease probabilities and larger overlaps of the disease probabilities with the maximum-
likelihood hypothesis (see also Appendix C); we find that already the D2S1 model works
considerably better than the D1S1 model for disease-disease interactions of relative strengths
K ab
i ≃ 0.3. A larger polarization means that we need a smaller number of observations
(medical tests) to obtain more definitive disease probabilities. A larger disease likelihood,
i /K a
15
here means that we are following the direction that is suggested by the most likely diseases.
In this sense, it appears that the two-sign/disease interactions could be very helpful in the
early stages of the diagnosis.
We checked that the above picture also holds if we start with different numbers of
observed signs, and if we double the magnitude of all the couplings. Similar behaviors
are observed also for larger problem sizes (see Appendix C). However, we see that for
i = −1,
ij ∈ (−2, +2)/√NS, the MP strategy gets closer to the random
strongly interacting models with much larger higher-order interactions, e.g., K 0
and K a,ab
i
∈ (−2, +2), K a,ab
strategy. The fact that the MP strategy does not work well in this strongly correlated regime
was indeed expected. Here, the greedy strategy is working better than the MPD strategy,
and both are still performing better than the random strategy.
IV. DISCUSSIONS
In summary, we showed that considering the sign-sign and disease-disease interactions can
significantly change the statistical importance of the signs and diseases. More importantly,
we found that these higher-order correlations could be very helpful in the process of diagnosis,
especially in the early stages of the diagnosis. The results in Figs. 3 and 4 (and similar
figures in appendices) also indicate the significance and relevance of optimization of the
diagnosis procedure, where a good strategy could considerably increase the polarization
and likelihood of the diseases compared to the random strategy. In addition, we devised
an approximate inference algorithm with a time complexity that grows linearly with the
number of interaction factors connecting the diseases to the signs, and exponentially with
the maximum number of signs that are associated with such interaction factors. For clarity,
in this work, we considered only algorithms of minimal structure and complexity. It would be
straightforward to employ more accurate learning and inference algorithms in constructing
the models and inferring the statistical information from the models. The challenge is,
of course, to go beyond the heuristic and greedy algorithms that we used to study the
multistage stochastic optimization problem of deciding on the relevance and order of the
medical observations.
In this study, we considered very simple structures for the prior probability of the dis-
eases P0(D) and the leak probability of the signs P (Snodisease). Obviously, depending
16
on the available statistical data, we can obtain more reliable models also for these prob-
ability distributions. Alternatively, we could employ the maximum entropy principle to
construct directly the joint probability distribution of the sign and disease variables using
the joint marginal probabilities P (Si, Sj; Da, Db). Note that, in practice, it is easier to ob-
tain this type of information than the stronger conditional probabilities P (Si, SjonlyDa)
and P (Si, SjonlyDa, Db). However, these measures are easier to model (or estimated by
experts), in the absence of enough observational data, because they present the sole effects
of single (or few) diseases.
The emphasize, in this study, was more on the diagnostic performances of the inferred
models than on the statistical significance of the selected models for a given set of clinical
data. A more accurate treatment of the model selection, for a finite collection of data,
accounts also the complexity of the models to avoid the over-fitting of the data. However,
we note that including the sign-sign and disease-disease interactions in the models is indeed
more natural than ignoring such correlations. Finally, to take into account the role of noises
in the model parameters, one should take the average of the objective function over the
probability distribution of the parameters, which is provided by the likelihood of the model
parameters.
Our proposed framework can be adapted to address assignment problems in cell biology,
immunology, and evolutionary biology [61 -- 63, 65].
In contrast to clinical problems, here
data availability might be less of a problem in near future. Advances in genomics, transcrip-
tomics, proteomics and metabolomics promise high resolution molecular characterization of
cells. Intensive research has also been directed towards live single cell analysis which allows
characterization of pathways from an initial state to a final state [64]. Our approach can
be used to do early assignments and thus not only provides accuracy but also an improved
sensitivity for diagnostics at the cellular level.
17
Appendix A: Details of the models
In this section, we give more details of the models, and show how the model parameters
are computed given the following true marginal probabilities:
Ptrue(Sinodisease),
Ptrue(Si, SjonlyDa),
Ptrue(Si, SjonlyDa, Db),
∀i,
∀i 6= j&a,
∀i 6= j&a 6= b.
(A1)
(A2)
(A3)
1. Case of one-disease-one-sign interactions: D1S1 model
The simplest model is obtained by considering only the one-disease factors besides the
leaky interactions. In addition, we assume that each factor contains only local sign fields,
i.e.,
φ0(S) = ePi K 0
i Si,
φa(SDa) = eDa Pi K a
i Si,
φab(SDa, Db) = · · · = 1.
Here, the couplings are
K 0
i =
K a
i =
1
2
1
2
Ptrue(Si = −1nodisease)(cid:19) ,
ln(cid:18) Ptrue(Si = +1nodisease)
Ptrue(Si = −1onlyDa)(cid:19) − K 0
ln(cid:18) Ptrue(Si = +1onlyDa)
i .
Then the partition function reads
Z(D) =Yi 2 cosh(K 0
i +Xa
DaK a
i )! .
The likelihood of diseases is given by
(A4)
(A5)
(A6)
(A7)
(A8)
P0(Da).
(A9)
L(DSo) =Yi∈O
e(K 0
i +Pa DaK a
i )Si
2 cosh(K 0
i +Pa DaK a
i )! ×Ya
Note that after summing over the sign variables the disease variables become correlated and
the problem of computing the marginals is not straightforward. The same problem arises in
computation of the probability distribution of unobserved signs,
M(SuSo) =XD Yi
e(K 0
i +Pa DaK a
i )Si
2 cosh(K 0
i +Pa DaK a
i )! ×Ya
18
P0(Da).
(A10)
For the moment, we assume the number of involved variables ND,S are small enough to
do the above computations exactly. For larger number of variables, we have to resort to
an approximate algorithm. We will describe such an approximate algorithm later when we
consider more serious examples.
2. Case of two-disease-one-sign interactions: D2S1 model
Now we consider the more interesting case of one- and two-disease interaction factors.
Again, we assume that each factor contains only local sign fields, i.e.,
φ0(S) = ePi K 0
i Si,
φa(SDa) = eDa Pi K a
i Si,
φab(SDa, Db) = eDaDb Pi K ab
i Si,
φabc(SDa, Db, Dc) = · · · = 1.
Here, the couplings are
K 0
i =
K a
i =
K ab
i =
1
2
1
2
1
2
Ptrue(Si = −1nodisease)(cid:19) ,
ln(cid:18) Ptrue(Si = +1nodisease)
Ptrue(Si = −1onlyDa)(cid:19) − K 0
ln(cid:18) Ptrue(Si = +1onlyDa)
Ptrue(Si = −1onlyDa, Db)(cid:19) − K 0
ln(cid:18) Ptrue(Si = +1onlyDa, Db)
i ,
i − K a
i − K b
i .
Still, we can easily obtain the partition function
Z(D) =Yi 2 cosh(K 0
i +Xa
DaK a
i +Xa<b
DaDbK ab
i )! .
(A11)
(A12)
(A13)
(A14)
(A15)
(A16)
(A17)
The likelihood reads
L(DSo) =Yi∈O
Similarly,
e(K 0
2 cosh(K 0
i +Pa DaK a
i +Pa<b DaDbK ab
i )Si
i +Pa DaK a
i +Pa<b DaDbK ab
i )! ×Ya
P0(Da).
(A18)
M(SuSo) =XD Yi
e(K 0
2 cosh(K 0
i +Pa DaK a
i +Pa<b DaDbK ab
i )Si
i +Pa DaK a
i +Pa<b DaDbK ab
19
i )! ×Ya
P0(Da).
(A19)
3.
Introducing the two-sign interactions: D1S2 and D2S2 models
A more challenging model is obtained with the two-sign interactions
φ0(S) = ePi K 0
i Si,
φa(SDa) = eDa[Pi K a
i Si+Pi<j K a
ijSiSj ],
φab(SDa, Db) = eDaDb[Pi K ab
i Si+Pi<j K ab
ij SiSj],
(A20)
(A21)
(A22)
φabc(SDa, Db, Dc) = · · · = 1.
Here, even computing the partition function is difficult, and from the beginning we have to
resort to approximations.
First of all, we need to obtain the couplings given the conditional probabilities,
P0(Si) ≡ Ptrue(Sinodisease),
Pa(Si, Sj) ≡ Ptrue(Si, SjonlyDa),
Pab(Si, Sj) ≡ Ptrue(Si, SjonlyDa, Db),
Pa(Si) ≡ Ptrue(SionlyDa),
Pab(Si) ≡ Ptrue(SionlyDa, Db).
As before, the zero-order couplings are given by
K 0
i =
1
2
P0(Si = −1)(cid:19) .
ln(cid:18) P0(Si = +1)
(A23)
(A24)
(A25)
(A26)
To obtain the other couplings, we resort to the so-called independent-pair approximation
[43], where we assume
Pa(Si, Sj) ∝ eK a
Pab(Si, Sj) ∝ eK ab
ijSiSj +ha
ijSi+ha
jiSj ,
ij SiSj +hab
ij Si+hab
ji Sj ,
i Si
Pa(Si) ∝ eha
Pab(Si) ∝ ehab
i Si.
In this way we obtain,
K a
ij =
K ab
ij =
1
4
1
4
Pa(Si = +1, Sj = −1)Pa(Si = −1, Sj = +1)(cid:19) ,
ln(cid:18)Pa(Si = +1, Sj = +1)Pa(Si = −1, Sj = −1)
Pab(Si = +1, Sj = −1)Pab(Si = −1, Sj = +1)(cid:19) − K a
ln(cid:18)Pab(Si = +1, Sj = +1)Pab(Si = −1, Sj = −1)
ij − K b
ij,
and
K a
i =
1
4Xj6=i
ln(cid:18) Pa(Si = +1, Sj = +1)Pa(Si = +1, Sj = −1)
Pa(Si = −1, Sj = −1)Pa(Si = −1, Sj = +1)(cid:19)
(NS − 2)
2
−
20
ln(cid:18) Pa(Si = +1)
Pa(Si = −1)(cid:19) − K 0
i ,
(A27)
(A28)
(A29)
(A30)
(A31)
i
g
n
g
n
a
h
c
f
o
y
t
i
l
i
b
a
b
o
r
p
i
e
s
a
e
s
d
P
M
e
h
t
i
g
n
g
n
a
h
c
f
o
y
t
i
l
i
b
a
b
o
r
p
i
n
g
s
P
M
e
h
t
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
(a)
D1S1
D2S1
D1S2
D2S2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
noise strength
(b)
D1S1
D2S1
D1S2
D2S2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
noise strength
FIG. 5. Probability of changing the most probable (MP) disease (a) and sign (b) variable in the
one-disease-one-sign (D1S1), two-disease-one-sign (D2S1), one-disease-two-sign (D1S2), and two-
disease-two-sign (D2S2) models. There are ND = 5 diseases, NS = 15 signs, and NO = 3 observed
signs with positive values. The couplings in the interaction factors are iid random numbers dis-
tributed uniformly in the specified intervals: K 0
i = −1, K a,ab
i
∈ (−1, +1), K a,ab
ij ∈ (−1, +1)/√NS.
ij ∈ (−ǫ, +ǫ)/√NS. The
A noise of strength ǫ changes the couplings by δK a,ab
i
∈ (−ǫ, +ǫ), δK a,ab
data have been obtained from 1000 independent realizations of the model parameters.
K ab
i =
1
4Xj6=i
ln(cid:18) Pab(Si = +1, Sj = +1)Pab(Si = +1, Sj = −1)
Pab(Si = −1, Sj = −1)Pab(Si = −1, Sj = +1)(cid:19)
ln(cid:18) Pab(Si = +1)
Pab(Si = −1)(cid:19) − K 0
(NS − 2)
−
2
i − K a
i − K b
i .
(A32)
One can find other approximation methods in the literature to address the above inverse
problem. For example, one can employ the exact fluctuation-response relations within the
Bethe approximation to obtain more accurate estimations for the marginal probability dis-
tribution of sign variables [36, 45].
21
Note that the model parameters are usually noisy because they may come from an ap-
proximate learning algorithm with a finite collection of clinical data. Thus, it is important
to know how the noise strength affects the marginal sign/disease probabilities.
In Fig.
5, we show how much the uncertainties in the model parameters are likely to change the
sign/disease with the highest probability of being positive/present. We see that this prob-
ability increases linearly with the relative strength of the noise in the model parameters
ǫ ≡ δK/K; for example, in a small model of ND = 5 diseases, NS = 15 signs with NO = 3
observed signs, a noise of strength ǫ = 0.1 changes the most probable sign/disease in nearly
%10 of the cases.
Appendix B: An approximate inference algorithm
In this section, we present an approximate way of computing the sign/disease marginal
probabilities. To be specific, we will focus on the two-disease-two-sign (D2S2) model.
We will consider the joint measure
P (S; D) = P (SD)P0(D) ∝ e−H(S;D),
(B1)
of the disease and sign variables. Here, we defined the energy function
H(S; D) ≡ −Xi
K 0
i Si −Xa
Xa<b
Da[Xi
DaDb[Xi
K a
ijSiSj]−
K a
i Si +Xi<j
i Si +Xi<j
K ab
K ab
ij SiSj] + ln Z(D) − ln P0(D).
(B2)
We may obtain the local probability marginals by a Monte Carlo (MC) algorithm for the
above system of interacting variables. But computing the above energy function for a given
configuration of the variables is computationally hard because computing Z(D) is in general
hard. Thus, to reduce the computation time we have to find a good approximation for the
partition function.
The exact expression for the partition function is
Z(D) =XS
φ0(S) ×Ya
φa(SDa) ×Ya<b
φab(SDa, Db),
(B3)
22
which in terms of the couplings reads
ePi hi(D)Si+Pi<j Jij(D)SiSj ,
DaDbK ab
i
,
Z(D) =XS
hi(D) ≡ K 0
Jij(D) ≡Xa
DaK a
i +Xa<b
i +Xa
DaK a
DaDbK ab
ij .
ij +Xa<b
(B4)
(B5)
(B6)
We are going to use the Bethe approximation to find an estimation for the above partition
function [34, 46, 47]. More precisely, in this approximation, we assume that the interaction
graph of the variables (defined by the couplings Jij) is a tree graph. We know how to
compute exactly and efficiently the local marginal probabilities and the partition function
in tree interaction graphs. The fact that there is no loop in a tree graph allows us to write
the marginal probability of a variable in terms of the cavity probabilities of the neighboring
variables, computed in the absence of the interested variable. These cavity probabilities are
the main quantities of the Bethe approximation. Here, e.g., we need to compute the cavity
marginals µi→j(Si), giving the probability of value Si for sign i in the absence of sign j. The
Bethe equations governing these cavity marginals are the following recursive equations [34]:
µi→j(Si) =
1
zi→j
ehiSi Yk6=i,j XSk
eJikSiSkµk→i(Sk)! ,
(B7)
where zi→j is a normalization constant. We can solve these equations by iteration starting
from random initial cavity marginals; the cavity marginals are updated according to the
above equations in a random and sequential way, until the algorithm converges. Then, we
use these stationary cavity marginals to compute the local probabilities and the partition
function [34].
In this way, for the partition function within the Bethe approximation, we obtain
Z(D) ≃ ePi ∆i(D)−Pi<j ∆ij(D),
XSj
e∆i(D) ≡XSi
e∆ij (D) ≡ XSi,Sj
ehiSiYj6=i
eJijSiSj µi→j(Si)µj→i(Sj).
eJij SiSj µj→i(Sj)
,
(B8)
(B9)
(B10)
Notice that here the cavity marginals µi→j(Si) depend implicitly on the disease variables.
Now, we can compute efficiently an approximate change in the energy function H(S; D)
23
when a disease variable is chosen for updating in a Monte Carlo algorithm. In the same way,
we may find an estimate of the likelihood of diseases by
L(DSo) =
Z(DSo)
Z(D)
P0(D),
where
Z(DSo) =XSu
φ0(S) ×Ya
φa(SDa) ×Ya<b
φab(SDa, Db),
(B11)
(B12)
can be estimated within the Bethe approximation, as we did for Z(D), but with fixed values
for the observed signs. This approximate expression for the likelihood is useful to find the
most likely diseases DM L maximizing the likelihood.
Finding the maximum likelihood hypothesis DM L, and obtaining good estimations for
the probability marginals of P (S; D) could be still very time consuming for a large number
of variables and frustrating energy functions.
In the following, we will utilize the Bethe
approximation to estimate the above marginal probabilities.
1. Belief propagation equations
The interaction factors associated to diseases a and (ab) are
K a
φa = exp Da[Xi
φab = exp DaDb[Xi
K a
i Si +Xi<j
i Si +Xi<j
K ab
ijSiSj]! ,
ij SiSj]! ,
K ab
(B13)
(B14)
respectively. Let us label the above interactions with α ∈ {a, b, . . . , (ab), (ac), . . .} and use
φα(SαDα) for the corresponding interaction factors. We represent the subset of sign and
disease variables in an interaction factor with ∂Sα and ∂Dα. The subset of interaction
factors depending on sign i and disease a are denoted by ∂i and ∂a, respectively. We will
assume that the resulting interaction graph is sparse, that is the numbers of interaction
factors (Ma, Mab) are small compared to NS, and the numbers of sign variables involved in
the interaction factors (ka, kab) are of order one. We take νi→α(Si) for the probability of
having value Si for sign i in the absence of interaction factor α. Similarly, define the cavity
probabilities νa→α(Da). In the Bethe approximation, we write approximate expressions for
24
these cavity marginals in terms of the messages received from the other interaction factors
in the system [34, 47]. More precisely,
Ψβ→i(Si),
i Si Yβ∈∂i6=α
νi→α(Si) ∝ eK 0
νa→α(Da) ∝ P0(Da) Yβ∈∂a6=α
Ψβ→a(Da)ΨZ→a(Da).
(B15)
(B16)
Note that the equation for νi→α(Si) gives the cavity marginal for an unobserved sign; the
cavity marginals of observed signs are fixed to values determined by the observations. The
cavity messages received from the interaction factors are given by
Ψα→i(Si) ∝ X{Da:a∈∂Dα} X{Sj:j∈∂Sα\i}
Ψα→a(Da) ∝ X{Db:b∈∂Dα\a} X{Sj :j∈∂Sα}
φα(SαDα) Ya∈∂Dα
φα(SαDα) Yb∈∂Dα\a
νa→α(Da) Yj∈∂Sα\i
νb→α(Db) Yj∈∂Sα
νj→α(Sj),
(B17)
νj→α(Sj).
(B18)
It remains to write the cavity marginals related to the partition function Z(D),
Ψα→a(Da),
νa→Z(Da) ∝ P0(Da) Yα∈∂a
ΨZ→a(Da) ∝ X{Db:b6=a}
e− ln Z(D)Yb6=a
νb→Z(Db).
(B19)
(B20)
The main difficulty comes from the last equation, where one has to sum over all possible
values of the disease variables and compute Z(D) for each configuration. We already know
how to compute the partition function by the Bethe approximation. But, the sum over the
disease variables grows exponentially with the number of variables. A naive approximation
to get around this complexity is obtained by a Bethe factorization of the partition function;
this is a nonnegative function of the disease variables, therefore, up to a normalization
constant, it can be interpreted as a probability measure of the associated variables. Suppose
µa(Da) and µα(Dα) are the local probability marginals of this measure. Then we may
approximate the partition function (up to a constant multiplication factor) by
Z(D) ∝Ya
µa(Da)Yα
µα(Dα)
Qa∈∂Dα µa(Da)! .
(B21)
Note that, we can obtain approximate expressions for the above local marginals by solving
the above equations without considering the interactions induced by the partition function
25
y
t
i
l
i
b
a
b
o
r
P
y
t
i
l
i
b
a
b
o
r
P
(a)
1
0.8
0.6
0.4
0.2
0
0
50
100
Sign
150
200
1
0.8
0.6
0.4
0.2
0
(b)
0 2 4 6 8 10 12 14 16 18 20
Disease
FIG. 6. The sign (a) and disease (b) marginal probabilities computed by the approximate belief
propagation algorithm for a single instance of the two-disease-two-sign (D2S2) model. The pa-
rameters are ND = 20, NS = 200, NO = 10, Ma = 20, Mab = 40, ka = 6, kab = 3. Each factor
includes all the possible two-sign interactions in addition to the local sign fields. The couplings are
iid random numbers distributed uniformly in the specified intervals: K 0
and K a
ij ∈ (−1, +1)/√ka, K ab
∈ (−1, +1),
ij ∈ (−1, +1)/√kab. The sign and disease variables in each factor are
i = −1, K a,ab
i
chosen randomly with a uniform probability.
Z(D). Moreover, the above approximation can be improved by considering the higher-order
marginals in the above factorization.
Now that Z(D) is factorized, we do not need any longer the cavity messages introduced
in Eqs. B19 and B20.
Instead, we will absorb the factors in the interaction factors φα,
26
working with the modified interactions
φa
,
µa(Da)
φa ≡
φab ≡ µa(Da)µb(Db)
φab
µab(Da, Db)
.
(B22)
(B23)
In summary, within the above approximation of the partition function, we first solve the
following equations for the cavity marginals, considering the modified interactions,
i Si Yβ∈∂i6=α
νi→α(Si) ∝ eK 0
νa→α(Da) ∝ P0(Da) Yβ∈∂a6=α
Ψβ→i(Si),
Ψβ→a(Da),
(B24)
(B25)
and
Ψα→i(Si) ∝ X{Da:a∈∂Dα} X{Sj:j∈∂Sα\i}
Ψα→a(Da) ∝ X{Db:b∈∂Dα\a} X{Sj :j∈∂Sα}
φα(SαDα) Ya∈∂Dα
φα(SαDα) Yb∈∂Dα\a
νa→α(Da) Yj∈∂Sα\i
νb→α(Db) Yj∈∂Sα
νj→α(Sj),
(B26)
νj→α(Sj).
(B27)
These equations are solved by iteration; we start from random initial messages, and update
the cavity marginals according to the above equations until the algorithm converges. Then,
we obtain the sign and disease marginal probabilities from similar equations, but taking into
account all the neighboring interactions, that is,
i Si Yα∈∂i
P (Si) ∝ eK 0
P (Da) ∝ P0(Da) Yα∈∂a
Ψα→i(Si),
Ψα→a(Da).
(B28)
(B29)
Figure 6 shows the sign and disease probabilities we obtain for the two-disease-two-sign
(D2S2) model with ND = 20 diseases, NS = 200 signs, and NO = 10 observed signs of
positive values. Here, we take a sparse interaction graph with Ma = 20 one-disease factors
and Mab = 40 two-disease factors connected randomly to ka = 6 and kab = 3 sign variables,
respectively. The quality of our approximation depends very much on the structure of
the interaction factors and strength of the associated couplings in the models. The Bethe
approximation is exact for interaction graphs that have a tree structure. It is expected to
work very well also in sparsely connected graphs, where the number of interaction factors
27
r
o
r
r
e
e
v
i
t
l
a
e
R
y
t
i
l
i
b
a
b
o
r
p
e
s
a
e
s
d
i
n
i
r
o
r
r
e
e
v
i
t
l
a
e
R
y
t
i
l
i
b
a
b
o
r
p
n
g
s
i
n
i
0.24
0.2
0.16
0.12
0.08
0.04
0
0.16
0.12
0.08
0.04
0
(a)
1
2
3
4
5
ka
6
7
8
9
(b)
1
2
3
4
5
ka
6
7
8
9
FIG. 7. Relative errors of the disease (a) and sign (b) marginal probabilities computed by the
approximate belief propagation algorithm for the two-disease-two-sign (D2S2) model. The pa-
rameters are ND = 5, NS = 20, NO = 3, Ma = 5, kab = 2, with Mab and ka taking different
values. Each factor includes all the possible two-sign interactions in addition to the local sign
fields. The couplings are iid random numbers distributed uniformly in the specified intervals:
i = −1, K a,ab
K 0
variables in each factor are chosen randomly with a uniform probability.
ij ∈ (−1, +1)/√kab. The sign and disease
ij ∈ (−1, +1)/√ka, K ab
∈ (−1, +1), and K a
i
(Ma, Mab) and the number of signs associated to an interaction factor (ka, kab) are small
compared to the total number of sign variables. This behavior is observed in Fig. 7, which
displays the relative errors in the marginal probabilities of sign/disease variables, obtained
by the above approximate algorithm. The time complexity of our approximate inference
algorithm grows linearly with the number of interaction factors, and exponentially with
the number of variables involved in such interactions; with ND = 500, NS = 5000, Ma =
500, Mab = 1000, ka = 10, kab = 5 the algorithm takes about one minute of CPU time on a
standard PC to compute the local marginals.
28
Appendix C: Comparing the diagnostic performance of the models
As explained in the main text, we reduce the multistage problem of diagnosis to a sequence
of two-stage problems. At each step, we use a greedy or heuristic strategy to choose the next
sign for observation. We study two different cases depending on the values that are assigned
to the observed signs: (I) in a real observation we find the true value of the observed sign,
(II) in a simulation of the observation process, the value of observed sign j is sampled from
its marginal probability P (Sj). In this section, we present more details of the two cases.
1. Case I
Here, we need to know the true sign values, therefore, we choose a true model with a clear
association of symptoms to diseases. As the benchmark, we take the exponential conditional
probability,
Ptrue(SD) =
1
Ztrue(D)
e−H(S,S∗(D)),
(C1)
where S∗(D) gives the symptoms of hypothesis D. We choose the symptoms randomly and
i)2/4 is the
Hamming distance (number of different signs) of the two sign configurations. As before,
uniformly from the space of sign configurations. Here H(S, S′) = PNS
i=1(Si − S′
we will take a factorized and unbiased prior disease probability P0(D). Note that there is
no sign-sign interaction in the above true model. Therefore, we need to consider only the
effects of disease-disease interactions, and study the D1S1 and D2S1 models. Let us assume
that we are given the true conditional marginals Ptrue(Sinodisease), Ptrue(SionlyDa), and
Ptrue(SionlyDa, Db). From these information, we can exactly obtain the model parameters
i , K a
K 0
Let us take a single realization of the true model for a small number of sign/disease
i as described in Appendix A.
i , and K ab
variables. This allows us to compute exactly the conditional marginal probabilities P (Si =
+1Da = 1) and P (Si = −1Da = 0), which are known in the literature as sensitivity and
specificity, respectively. As Fig. 8 shows, both the probabilities are in average closer to the
true values in the D2S1 model.
Now, we consider a sequential process of T decisions and real observations; at each step
an unobserved sign is chosen for observation and the observed sign takes the true value, given
by the symptoms S∗(D) of the true hypothesis D. This hypothesis is selected randomly and
29
y
t
i
c
i
f
i
c
e
p
s
y
t
i
v
i
t
i
s
n
e
s
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
True
D1S1
D2S1
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
sign
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
sign
FIG. 8. Sensitivity, P (Si = +1Da = 1), and specificity, P (Si = −1Da = 0), in the D1S1 and
D2S1 models for the exponential benchmark model. The data are for one of the diseases in a single
realization of the true model. The model parameters are obtained from the conditional marginals
of the exponential true model. Here we have ND = 5 diseases, NS = 20 signs.
uniformly at the beginning of the process. At each step, we compute the sign marginals
P (Sj) in the constructed models, and use the MP (or greedy) strategy to choose the next
observation. Figure 9 shows the RMS errors ∆D and ∆S in the marginal disease and sign
probabilities, respectively, during such a process. Here, we are using the MP strategy and
the errors are defined as follows,
∆D ≡ 1
∆S ≡ 1
ND Xa
NU Xi∈U
(P (Da = 1) − Ptrue(Da = 1))2!1/2
(P (Si = 1) − Ptrue(Si = 1))2!1/2
,
,
(C2)
(C3)
where U is the subset of unobserved signs with NU = U. Figure 10 displays the inferred
disease probabilities P (Da = 1) in the same process for cases in which only one or two
30
0.6
y
t
i
l
i
0.5
D1S1
D2S1
(a)
r
o
r
r
e
b
a
b
o
r
p
e
s
a
e
s
d
n
i
i
y
t
i
l
i
r
o
r
r
e
b
a
b
o
r
p
n
g
s
n
i
0.4
0.3
0.2
0.1
0
2
6
8
4
number of observations
10 12 14 16
D1S1
D2S1
(b)
0.25
0.2
0.15
i
0.1
0
2
6
8 10 12 14 16
4
number of observations
FIG. 9. The RMS error in disease (a) and sign (b) marginal probabilities in the D1S1 and
D2S1 models for the exponential benchmark model. The model parameters are obtained from
the conditional marginals of the true model. There are ND = 5 diseases, NS = 20 signs, and
the algorithm starts with NO = 3 observed signs for a randomly selected hypothesis D. An
unobserved sign is chosen for observation by the MP strategy using the inferred sign probabilities,
and the observed sign takes the true value given by S∗(D). The data are results of averaging over
1000 independent realizations of the true model and the observation process.
diseases are present. It is seen that both the D1S1 and D2S1 models end up with the right
diagnosis when only one diseases is present (D1 = 1, D2 = D3 = · · · = 0). But, the D1S1
model fails when there are two diseases in the hypothesis (D1 = D2 = 1, D3 = D4 = · · · = 0),
whereas the other model can still find the right diseases. However, we find that even the
D2S1 model gets confused when there are more than two diseases in the hypothesis. Then,
we would need to consider more sophisticated models with interactions involving more than
two diseases.
To see the importance of sign-sign interactions, we consider another benchmark model
31
(b) D2S1: only D1
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
number of observations
(d) D2S1: only D1,D2
(a) D1S1: only D1
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
number of observations
(c) D1S1: only D1,D2
5
4
e
s
a
3
e
s
d
2
i
1
5
4
e
s
a
3
e
s
d
2
i
1
1
0.8
0.6
0.4
0.2
0
0.4
0.35
0.3
0.25
0.2
5
4
e
s
a
3
e
s
d
2
i
1
5
4
e
s
a
3
e
s
d
2
i
1
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
number of observations
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
number of observations
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
FIG. 10. The inferred disease probabilities vs the number of observations for the exponential
benchmark model. The data are for the cases in which only one (a,b) or two (c,d) diseases are
present. The model parameters are obtained from the conditional marginals of the true model.
There are ND = 5 diseases, NS = 20 signs, and the algorithm starts with NO = 3 observed signs
for a randomly selected hypothesis D. An unobserved sign is chosen for observation by the MP
strategy using the inferred sign probabilities. The observed sign takes the true value given by
S∗(D). The data are results of averaging over 1000 independent realizations of the true model and
the observation process.
given by the following conditional probability:
Ptrue(SD) =
1
Ztrue(D)(cid:18)
1
1 + H(S, S∗(D))(cid:19) .
(C4)
Given a disease hypothesis D and the associated symptoms S∗(D), we start from a few
number of randomly chosen observed sings from the set of symptoms S∗(D). At each step,
we compute the inferred sign probabilities P (Si) and use the greedy algorithm to choose an
unobserved sign for observation. The observed sign at each time step takes the true value
given by S∗(D). Figure 11 shows the overlap of the inferred probabilities P (Da = 1) with
the true hypothesis for cases in which only one or two diseases are present. For comparison,
32
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
(a)
D1S1 (G)
D2S1 (G)
D2S2 (G)
D2S2 (Random)
0.1
0.08
0.06
0.04
0.02
0
0
2
6
4
number of observations
8
10 12 14 16
(b)
D1S1 (G)
D2S1 (G)
D2S2 (G)
D2S2 (Random)
0.1
0.08
0.06
0.04
0.02
0
0
2
6
4
number of observations
8
10 12 14 16
FIG. 11. Overlap of the inferred disease marginals with the true hypothesis for the power-law
benchmark models. The data are for the cases in which only one (a) or two (b) diseases are
present. The model parameters are obtained from the conditional marginals of the true model.
There are ND = 5 diseases, NS = 20 signs, and the algorithm starts with NO = 3 observed signs
for a randomly selected hypothesis D. An unobserved sign is chosen for observation by the greedy
(G) strategy using the inferred sign probabilities. The observed sign takes the true value given by
S∗(D). The data are results of averaging over 1000 independent realizations of the true model and
the observation process.
we also show the results obtained by the random strategy. The figure shows that we can
obtain higher overlaps between the inferred disease marginals and the true hypothesis by
considering the sign-sign interactions. Nevertheless, we find that even the D2S2 model fails
in cases with more than two diseases.
33
2. Case II
Here, the model parameters are chosen randomly and uniformly in symmetric intervals
of real numbers; e.g., K a,ab
ij ∈ (−1, +1)/√NS. The leaky couplings are
i = −1, corresponding to small sign probabilities P (Si = 1nodisease) ≃ 0.05.
We assume that all the possible one-disease and two-disease interaction factors are present
∈ (−1, +1), K a,ab
set to K 0
i
in the models. Moreover, inside each factor we have all the possible two-sign interactions
in addition to the local sign fields; that is why we scale the two-sign couplings K a,ab
ij with
the inverse square root of the number of sign variables. The prior disease probabilities are
uniform P0(Da = 1) = 0.5.
We start from a small number of randomly selected observed signs with random values.
Then, at each step, we choose an unobserved sign for observation with stochastic findings
sampled from the associated probability distributions. The process is repeated 1000 times for
different realizations of the couplings and findings of the observations to obtain the average
polarization and likelihood of diseases. For reference, we also show the results that were
obtained by a random strategy, where an unobserved sign is chosen randomly and uniformly
for observation.
In Fig. 12 we see that the average performance of the greedy strategy
changes from that of the MPD strategy for small number of observations to that of the MP
strategy for larger number of observations.
However, as Fig. 13 shows, for strongly correlated models with much larger higher-order
interactions, e.g., K 0
i = −1 and K a,ab
i
∈ (−2, +2), K a,ab
ij ∈ (−2, +2)/√NS, the difference
between the performances of the most positive (MP) strategy and the random strategy
diminishes. Furthermore, here the greedy strategy performs better than the MP and MPD
strategies. The fact that such a heuristic algorithm does not work well in this regime was
indeed expected; it is this type of finding that makes strongly interacting systems interesting,
and in need for more elaborate methods and algorithms.
Finally, Fig. 14 displays the average polarization of disease probabilities obtained by
the approximate inference algorithm (BP) for much larger number of sign/disease variables.
Here, the maximum likelihood hypothesis (DM L in the MPD strategy) is simply approxi-
mated by the most probable diseases suggested by the marginal disease probabilities. That
34
n
o
i
t
l
a
z
i
r
a
o
p
e
s
a
e
s
d
i
n
o
i
t
l
a
z
i
r
a
o
p
e
s
a
e
s
d
i
l
n
o
i
t
a
z
i
r
a
o
p
e
s
a
e
s
d
i
0.4
0.38
0.36
0.34
0.32
0.3
0.28
0.48
0.45
0.42
0.39
0.36
0.33
0.3
0.45
0.42
0.39
0.36
0.33
0.3
(a)
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
D1S1: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
(c)
D2S1: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
(e)
D1S2: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
0.38
0.36
0.34
0.32
0.3
0.28
0.26
0.24
0.46
0.44
0.42
0.4
0.38
0.36
0.34
0.32
0.3
0.28
0.26
0.44
0.42
0.4
0.38
0.36
0.34
0.32
0.3
0.28
0.26
(b)
D1S1: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
(d)
D2S1: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
(f)
D1S2: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
FIG. 12. Diagnostic performance of the models vs the number of observations for different strate-
gies. The exact disease-polarization (DP) and disease-likelihood (DL) obtained by the greedy,
MP, MPD, and random strategies in the one-disease-one-sign (D1S1, (a)-(b)), two-disease-one-
sign (D2S1, (c)-(d)), and one-disease-two-sign (D1S2, (e)-(f)) models. There are ND = 5 dis-
eases, NS = 20 signs, and the algorithm starts with NO = 4 observed signs with positive values.
The coupling constants are iid random numbers distributed uniformly in the specified intervals:
i = −1, K a,ab
K 0
independent realizations of the model parameters.
ij ∈ (−1, +1)/√NS. The data are results of averaging over 1000
∈ (−1, +1), K a,ab
i
35
n
o
i
t
l
a
z
i
r
a
o
p
e
s
a
e
s
d
i
n
o
i
t
a
z
i
r
a
o
p
l
e
s
a
e
s
d
i
0.5
0.45
0.4
0.35
0.5
0.48
0.46
0.44
0.42
0.4
0.38
0.36
(a)
D1S1
D2S1
D1S2
D2S2
0
2
4
6
8
10 12 14 16
number of observations
(c)
D2S2: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
d
o
o
h
i
l
e
k
i
l
e
s
a
e
s
d
i
0.5
0.45
0.4
0.35
0.3
0.5
0.48
0.46
0.44
0.42
0.4
0.38
0.36
0.34
(b)
D1S1
D2S1
D1S2
D2S2
0
2
4
6
8
10 12 14 16
number of observations
(d)
D2S2: MPD
G
MP
Random
0
2
4
6
8
10 12 14 16
number of observations
FIG. 13. Diagnostic performance of the models vs the number of observations for larger couplings.
(top) The exact disease polarization (a) and disease likelihood (b) obtained by the MP strategy
in the one-disease-one-sign (D1S1), two-disease-one-sign (D2S1), one-disease-two-sign (D1S2),
and two-disease-two-sign (D2S2) models. There are ND = 5 diseases, NS = 20 signs, and the
algorithm starts with NO = 4 observed signs with positive values. The couplings in the interaction
i = −1, K a,ab
∈
ij ∈ (−2, +2)/√NS. (bottom) Comparing the exact polarization (c) and likelihood
factors are iid random numbers distributed uniformly in the specified intervals: K 0
(−2, +2), K a,ab
(d) of the diseases obtained by the greedy, MP, MPD, and random strategies, for the D2S2 model.
i
The data are results of averaging over 1000 independent realizations of the model parameters.
is, we take Da = 0 if P (Da = 0) > 1/2, otherwise Da = 1.
[1] Merrell, A.J. and Stanger, B.Z. (2016) Adult cell plasticity in vivo: de-differentiation and
transdifferentiation are back in style. Nature Review Molecular Cell Biology, 17(7), 413-425.
[2] DuPage, M., and Bluestone, J.A. (2016). Harnessing the plasticity of CD4+ T cells to treat
immune-mediated disease. Nature Review Immunology, 16(3), 149-163.
36
[3] Holzel, M., Bovier, A., and Tuting, T. (2016). Plasticity of tumour and immune cells: a source
of heterogeneity and a cause for therapy resistance? Nature Review Cancer, 13(5), 365-376.
[4] Saeys, Y., Van Gassen, S., and Lambrecht, B.N. (2016). Computational flow cytometry: help-
ing to make sense of high-dimensional immunology data. Nature Reviews Immunology 16,
449462.
[5] Patti, G.j., Yanes, O., Siuzdak, G. (2012). Metabolomics: the apogee of the omics, trilogy.
Nature Reviews Molecular Cell Biology 13, 263-269.
[6] Kasper, D., Fauci, F., Hauser, S., Longo, D., Jameson, J., Loscalzo, J. (2015). Harrison's Prin-
ciples of Internal Medicine. 19 edition, McGraw-Hill Education, 3000 pages, ISBN: 0071802150.
[7] Papadakis, M., McPhee, S.J., Rabow, M.W. (2016). Current Medical Diagnosis and Treat-
ment. 55 edition, LANGE CURRENT Series, 1920 pages, ISBN: 0071845097.
[8] Lerner, A. J. (2006). Diagnostic gold standards or consensus diagnostic criteria. Diagnostic
criteria in neurology, Humana Press.
[9] Coggon, D., Martyn, C., Palmer, K.T., Evanoff, B. (2005). Assessing case definitions in the
absence of a diagnostic gold standard. International Journal of Epidemiology, 34(4), 949-52.
[10] Greenes, R. (2014). Clinical Decision Support, The Road to Broad Adoption. 2nd Edition,
Academic Press, 930 pages, ISBN: 9780123984760.
[11] Miller, R. A., Pople Jr, H. E., and Myers, J. D. (1982). Internist-I, an experimental computer-
based diagnostic consultant for general internal medicine. New England Journal of Medicine,
307(8), 468-476.
[12] Heckerman, D. E. (1987). Formalizing heuristic methods for reasoning with uncertainty. Tech-
nical Report KSL-88-07, Medical Computer Science Group, Section on Medical Informatics,
Stanford University, Stanford, CA.
[13] Horvitz, E. J., Breese, J. S., and Henrion, M. (1988). Decision theory in expert systems and
artificial intelligence. International journal of approximate reasoning, 2(3), 247-302.
[14] Shortliffe, E. H. (1993). The adolescence of AI in medicine: will the field come of age in
the'90s?. Artificial intelligence in medicine, 5(2), 93-106.
[15] Adams, J. B. (1976). A probability model of medical reasoning and the MYCIN model. Math-
ematical biosciences, 32(1), 177-186.
[16] Buchanan, B. G., and Shortliffe, E. H. (Eds.). (1984). Rule-based expert systems (Vol. 3).
Reading, MA: Addison-Wesley.
37
[17] Miller, R., Masarie, F. E., and Myers, J. D. (1985). Quick medical reference (QMR) for
diagnostic assistance. MD computing: computers in medical practice, 3(5), 34-48.
[18] Barnett, G. O., Cimino, J. J., Hupp, J. A., and Hoffer, E. P. (1987). DXplain: an evolving
diagnostic decision-support system. Jama, 258(1), 67-74.
[19] Spielgelharter, D. J. (1987). Probabilistic Expert Systems in Medicine. Statistical Science, 2,
3-44.
[20] Bankowitz, R. A., McNeil, M. A., Challinor, S. M., Parker, R. C., Kapoor, W. N., and Miller,
R. A. (1989). A computer-assisted medical diagnostic consultation service:
implementation
and prospective evaluation of a prototype. Annals of Internal Medicine, 110(10), 824-832.
[21] Heckerman D., A tractable inference algorithm for diagnosing multiple diseases, In Machine
Intelligence and Pattern Recognition: Uncertainty in artificial Intelligence 5, Henrion M.,
Shachter R., Kanal L. N., Lemmer J. F., eds. Amsterdam North Holland Publ. Comp. 1990:
163-172.
[22] Shwe, M. A., Middleton, B., Heckerman, D. E., Henrion, M., Horvitz, E. J., Lehmann, H. P.,
and Cooper, G. F. (1991). Probabilistic diagnosis using a reformulation of the INTERNIST-
1/QMR knowledge base. Methods of information in Medicine, 30(4), 241-255.
[23] Heckerman, D. E., and Shortliffe, E. H. (1992). From certainty factors to belief networks.
Artificial Intelligence in Medicine, 4(1), 35-52.
[24] Nikovski, D. (2000). Constructing Bayesian networks for medical diagnosis from incomplete
and partially correct statistics. Knowledge and Data Engineering, IEEE Transactions on,
12(4), 509-516.
[25] Przulj, N. and Malod-Dognin, N. (2016). Network analytics in the age of big data. Science
353(6295), 123-124.
[26] Hamaneh, M.B. and Yu, Y-K. (2015). DeCoaD: determining correlations among diseases using
protein interaction networks. BMC Research Notes, 8:226.
[27] Zitnik, M., Janjic, V., Larminie, C., Zupan, B., and Natasa Przuljb, N. (2013). Discovering
disease-disease associations by fusing systems-level molecular data. Sci Rep. 3: 3202.
[28] Liu, W., Wu, A., Pellegrini, M., and Wanga, X. (2015). Integrative analysis of human protein,
function and disease networks. Sci Rep. 2015; 5: 14344.
[29] Suratanee, A., and Plaimas, K. (2015). DDA: A Novel Network-Based Scoring Method to
Identify DiseaseDisease Associations. Bioinform Biol Insights. 9: 175186.
38
[30] Gustafsson, M., Nestor, C.E., Zhang, H., Barabasi, A.L., Baranzini, S., Brunak, S., Chung,
K.F. et al. (2014) Modules, networks and systems medicine for understanding disease and
aiding diagnosis. Genome Medicine 6:82.
[31] Liu, C-C., Tseng, Y-T., Li, W., Wu, C-Y., Mayzus, I., Rzhetsky, A., Sun, F., Waterman, M.,
et al. (2015). DiseaseConnect: a comprehensive web server for mechanism-based diseasedisease
connections. Nucl. Acids Res. 42 (W1): W137-W146.
[32] Sun, K., Goncalves, J.P., Larminie, C., and Przulj, N. (2014). Predicting disease associations
via biological network analysis. BMC Bioinformatics 15:304
[33] Jordan, M. I. (2004). Graphical models. Statistical Science, 140-155.
[34] Mezard, M., and Montanari, A. (2009). Information, physics, and computation. Oxford Uni-
versity Press.
[35] Altarelli, F., Braunstein, A., Ramezanpour, A., and Zecchina, R. (2011). Stochastic matching
problem. Physical review letters, 106(19), 190601.
[36] Ricci-Tersenghi, F. (2012). The Bethe approximation for solving the inverse Ising problem:
a comparison with other inference methods. Journal of Statistical Mechanics: Theory and
Experiment, 2012(08), P08015.
[37] Edwin, T. (2003). Jaynes. Probability theory: the logic of science. Cambridge university press,
10, 33.
[38] Kappen, H. J., and Rodriguez, F. B. (1998). Efficient learning in Boltzmann machines using
linear response theory. Neural Computation, 10(5), 1137-1156.
[39] Tanaka, T. (1998). Mean-field theory of Boltzmann machine learning. Physical Review E,
58(2), 2302.
[40] Schneidman, E., Berry, M. J., Segev, R., and Bialek, W. (2006). Weak pairwise correlations
imply strongly correlated network states in a neural population. Nature, 440(7087), 1007-1012.
[41] Cocco, S., Leibler, S., and Monasson, R. (2009). Neuronal couplings between retinal gan-
glion cells inferred by efficient inverse statistical physics methods. Proceedings of the National
Academy of Sciences, 106(33), 14058-14062.
[42] Weigt, M., White, R. A., Szurmant, H., Hoch, J. A., and Hwa, T. (2009). Identification of
direct residue contacts in proteinprotein interaction by message passing. Proceedings of the
National Academy of Sciences, 106(1), 67-72.
[43] Roudi, Y., Aurell, E., and Hertz, J. A. (2009). Statistical physics of pairwise probability
39
models. Frontiers in computational neuroscience, 3.
[44] Bailly-Bechet, M., Braunstein, A., Pagnani, A., Weigt, M., and Zecchina, R. (2010). Infer-
ence of sparse combinatorial-control networks from gene-expression data: a message passing
approach. BMC bioinformatics, 11(1), 355.
[45] Nguyen, H. C., and Berg, J. (2012). Mean-field theory for the inverse Ising problem at low
temperatures. Physical review letters, 109(5), 050602.
[46] Pearl, J. (2014). Probabilistic reasoning in intelligent systems: networks of plausible inference.
Morgan Kaufmann.
[47] Kschischang, F. R., Frey, B. J., and Loeliger, H. A. (2001). Factor graphs and the sum-product
algorithm. Information Theory, IEEE Transactions on, 47(2), 498-519.
[48] Mzard, M., Parisi, G., and Zecchina, R. (2002). Analytic and algorithmic solution of random
satisfiability problems. Science, 297(5582), 812-815.
[49] Yedidia, J. S., Freeman, W. T., and Weiss, Y. (2003). Understanding belief propagation and
its generalizations. Exploring artificial intelligence in the new millennium, 8, 236-239.
[50] Birge, J. R., and Louveaux, F. (2011). Introduction to stochastic programming. Springer
Science & Business Media.
[51] Kleywegt, A. J., Shapiro, A., and Homem-de-Mello, T. (2002). The sample average approx-
imation method for stochastic discrete optimization. SIAM Journal on Optimization, 12(2),
479-502.
[52] Heyman, D. P., and Sobel, M. J. (2003). Stochastic Models in Operations Research: Stochastic
Optimization (Vol. 2). Courier Corporation.
[53] Garey, M. R., and Johnson, D. S. (1979). Computers and intractability: A guide to the theory
of NP-completeness. W.H. Freeman.
[54] Cooper, G. F. (1990). The computational complexity of probabilistic inference using Bayesian
belief networks. Artificial intelligence, 42(2), 393-405.
[55] Henrion, M. (1990), Towards efficient probabilistic diagnosis in multiply connected belief net-
works. Oliver, R. M., Smith, J. Q., eds. Influence Diagrams, Belief Nets and Decision Analysis.
Chichester: Wiley, 385-407.
[56] Heckerman, D., Geiger, D., and Chickering, D. M. (1995). Learning Bayesian networks: The
combination of knowledge and statistical data. Machine learning, 20(3), 197-243.
[57] Chickering, D. M. (1996). Learning Bayesian networks is NP-complete. In Learning from data
40
(pp. 121-130). Springer New York.
[58] Lezon, T. R., Banavar, J. R., Cieplak, M., Maritan, A., and Fedoroff, N. V. (2006). Using the
principle of entropy maximization to infer genetic interaction networks from gene expression
patterns. Proceedings of the National Academy of Sciences, 103(50), 19033-19038.
[59] Bialek, W., and Ranganathan, R. (2007). Rediscovering the power of pairwise interactions.
arXiv preprint arXiv:0712.4397.
[60] Banavar, J. R., Maritan, A., and Volkov, I. (2010). Applications of the principle of maximum
entropy: from physics to ecology. Journal of Physics: Condensed Matter, 22(6), 063101.
[61] Barabsi, A.L. and Oltvai Z.A. (2004) Network biology: understanding the cell's functional
organization. Nature Reviews Genetics 5, 101-113.
[62] Kim, M., Rai, N., Zorraquino, V., and Tagkopoulos, I. (2016) Multi-omics integration accu-
rately predicts cellular state in unexplored conditions for Escherichia coli. Nature Communi-
cations 7,13090.
[63] Ebrahim, A., Brunk, E., Tan, J., O'Brien, E.J., Kim, D., Szubin, R., Lerman, J.A., Lechner,
A., Sastry, A., Bordbar, A., Feist, A.M., and Palsson, B.O. (2016) Multi-omic data integration
enables discovery of hidden biological regularities. Nature Communications 7, 13091.
[64] Spiller, D.J., Wood, C.D., Rand, D.A., and White, M.R.H. (2010) Measurement of single-cell
dynamics. Nature 465, 736745.
[65] Candia, J., Maunu, R., Driscoll, M., Biancotto, A., Dagur, P., et al. (2013) From Cellular
Characteristics to Disease Diagnosis: Uncovering Phenotypes with Supercells. PLOS Compu-
tational Biology 9(9): e1003215.
41
l
n
o
i
t
a
z
i
r
a
o
p
e
s
a
e
s
d
i
l
n
o
i
t
a
z
i
r
a
o
p
e
s
a
e
s
d
i
(a)
0.5
0.4
0.3
0.2
0.1
0
0
D1S1
D2S1
D1S2
D2S2
100
200
300
400
500
number of observations
0.5
0.45
0.4
0.35
0.3
0.25
D2S2: MPD
MP
Random
(b)
0
100
200
300
400
500
number of observations
FIG. 14. Diagnostic performance of the models vs the number of observations for a larger number
of sign/disease variables.
(a) The approximate disease-polarization (DP) obtained by the MP
strategy in the one-disease-one-sign (D1S1), two-disease-one-sign (D2S1), one-disease-two-sign
(D1S2), and two-disease-two-sign (D2S2) models. Here we have ND = 200 diseases, NS = 2000
signs, and the algorithm starts with NO = 50 observed signs with positive values. The graph
parameters are: Ma = 200, Mab = 0, ka = 8 in the D1S1 and D1S2 models, and Ma = 200, Mab =
1000, ka = 8, kab = 4 in the D2S1 and D2S2 models. The interaction factors in the D1S2 and
D2S2 models include all the possible two-sign interactions in addition to the local sign fields. The
couplings in the interaction factors are iid random numbers distributed uniformly in the specified
intervals: K 0
i = −1, K a,ab
i
∈ (−1, +1), K a,ab
polarization of the diseases obtained by the MP, MPD, and random strategies, for the D2S2 model.
ij ∈ (−1, +1)/pka,ab. (b) Comparing the approximate
The data are results of averaging over 100 independent realizations of the model parameters.
42
|
1710.06914 | 1 | 1710 | 2017-09-26T20:05:50 | The Influence of Ceramide Tail Length on the Structure of Bilayers Composed of Stratum Corneum Lipids | [
"physics.bio-ph",
"cond-mat.soft",
"physics.chem-ph"
] | Lipid bilayers composed of non-alpha hydroxy sphingosine ceramide (CER NS), cholesterol (CHOL), and free fatty acids (FFA), which are components of the human skin barrier, are studied via molecular dynamics simulations. Since mixtures of these lipids exist in dense gel phases with little molecular mobility at physiological conditions, care must be taken to ensure that the simulations become decorrelated from the initial conditions. Thus, we propose and validate an equilibration protocol based on simulated tempering in which the simulation takes a random walk through temperature space, allowing the system to break out of metastable configurations and hence become decorrelated form its initial configuration. After validating the equilibration protocol, the effects of the lipid composition and ceramide tail length on bilayer properties are studied. Systems containing pure CER NS, CER NS + CHOL, and CER NS + CHOL + FFA, with the CER fatty acid tail length varied within each CER NS-CHOL-FFA composition, are simulated. The bilayer thickness is found to depend on the structure of the center of the bilayer, which arises as a result of the tail length asymmetry between the lipids studied. The hydrogen bonding between the lipid headgroups and with water is found to change with the overall lipid composition, but is mostly independent of the CER fatty acid tail length. Subtle differences in the lateral packing of the lipid tails are also found as a function of CER tail length. Overall, these results provide insight into the experimentally observed trend of altered barrier properties in skin systems where there are more ceramides with shorter tails present. | physics.bio-ph | physics | The Influence of Ceramide Tail Length on the
Structure of Bilayers Composed of Stratum
Corneum Lipids
T. C. Moore, R. Hartkamp, C. R. Iacovella, A. L. Bunge, and C. McCabe
Condensed Title: Stratum Corneum Lipid Bilayers
1
ABSTRACT
Lipid bilayers composed of non-alpha hydroxy sphingosine ceramide (CER NS), cholesterol
(CHOL), and free fatty acids (FFA), which are components of the human skin barrier, are studied
via molecular dynamics simulations. Since mixtures of these lipids exist in dense gel phases with
little molecular mobility at physiological conditions, care must be taken to ensure that the
simulations become decorrelated from the initial conditions. Thus, we propose and validate an
equilibration protocol based on simulated tempering in which the simulation takes a random
walk through temperature space, allowing the system to break out of metastable configurations
and hence become decorrelated form its initial configuration. After validating the equilibration
protocol, the effects of the lipid composition and ceramide tail length on bilayer properties are
studied. Systems containing pure CER NS, CER NS + CHOL, and CER NS + CHOL + FFA,
with the CER fatty acid tail length varied within each CER NS-CHOL-FFA composition, are
simulated. The bilayer thickness is found to depend on the structure of the center of the bilayer,
which arises as a result of the tail length asymmetry between the lipids studied. The hydrogen
bonding between the lipid headgroups and with water is found to change with the overall lipid
composition, but is mostly independent of the CER fatty acid tail length. Subtle differences in
the lateral packing of the lipid tails are also found as a function of CER tail length. Overall, these
results provide insight into the experimentally observed trend of altered barrier properties in skin
systems where there are more ceramides with shorter tails present.
2
INTRODUCTION
The stratum corneum (SC) layer of the skin acts as the main barrier against chemical penetrants
and water loss and is composed of a brick-and-mortar-like arrangement of corneocytes
surrounded by a dense, lamellar-structured extracellular lipid matrix. This matrix is a complex
mixture primarily composed of an equimolar ratio of ceramides (CERs), which are composed of
a sphingoid base linked to a saturated fatty acid chain (in most cases), cholesterol (CHOL), and
free fatty acids (FFAs) (1, 2). Since the extracellular matrix is widely thought to be the only
continuous path for permeation through the SC, proper lipid organization and composition of the
lamellar membrane is generally considered crucial for a fully functioning skin barrier. This is
supported by the fact that many skin diseases that are characterized by a reduced barrier function
also exhibit an altered lipid composition and organization compared to healthy skin (3-6). For
example, a decrease has been observed in the average number of CER carbon atoms in the SC of
patients who suffer from atopic eczema (3, 7) and, compared with controls, less dense lipid
packing and changes in x-ray scattering patterns (3). Of the two CER tails, reduced CER carbon
numbers appear to correspond with a shorter fatty acid (FA) tail rather than the sphingoid base
tail (8).
The complex nature of the human SC lipid matrix, in that it features 15 unique subclasses
of CER lipids and a distribution of CER and FFA chain lengths, has made it challenging to
establish the influence of individual lipid species on barrier properties (9). As a result, many
experiments have focused on model systems that are composed of simpler mixtures of synthetic
lipids, whose composition and chain lengths can be precisely controlled (10-21). For example,
considering studies of mixtures containing non-alpha hydroxyl sphingosine CER (NS), CHOL,
and FFA, similar to those considered in this work, Mojumdar et al. (14) showed that CHOL is
3
essential to the characteristic phase behavior, and hence barrier properties, of such mixtures, and
Školová et al. (19) found that CER NS-based membranes with short acyl chains had increased
permeability compared to long acyl chain CERs. However, while global structural properties of
SC lipid lamellae can be inferred from experiment (e.g., from spectroscopic, scattering, and
diffraction data), including information about the localization of molecules across the lamellae,
the underlying molecular-level details and in-plane morphology often remain elusive.
Furthermore, structural examination using neutron scattering is quite laborious, as many
independent systems, each with different deuterated lipids and/or concentrations of deuterated
lipids, must be considered to provide a single picture of the structure, and the availability of
deuterated lipids may limit the study of specific systems (15, 22-24).
Molecular simulation, which offers atomic-level resolution of the entire 3-dimensional
structure of the lipids at picosecond timescales, has been used to provide a more direct
understanding of the molecular interactions in CER-based membranes (25-38), with several
studies focusing on the dependence of membrane structure on lipid composition. For example,
Das et al. (26) found that CHOL compresses a CER NS bilayer membrane and increases the
interdigitation of CER tails in opposing bilayer leaflets. Gupta and Rai (31) simulated a series of
lipid bilayers composed of CER NS, CHOL, and FFA with varying compositions and found that
CHOL tends to sit away from the lipid-water interface and increases the thermal stability of the
bilayers. The effect of the CER tail length on bilayer properties has received considerably less
attention. Paloncyová et al. showed that extremely short CER NS FA tails (e.g., 6 carbons) cause
a disruption in the headgroup packing of CER bilayers, which gives rise to an increased
permeability (37); however, the CER NS FA tail lengths considered were much shorter than
those found in the SC. More recently, Gupta et al. showed that the permeability of water across
4
pure CER NS bilayers with more biologically relevant tail lengths decreases monotonically with
the CER FA tail length (30), although a structural explanation for this trend was not presented.
While these computational studies have provided important insight, they also highlight a
significant challenge in studying SC lipids with molecular simulation. Specifically, due to the
dense packing of the lipid tails, the CER-rich bilayer systems demonstrate negligible molecular
diffusion at physiologically relevant temperatures (i.e., 305 K). Even in simulations conducted at
elevated temperatures, e.g., 340 K, studies have demonstrated minimal in-plane diffusion of the
lipids, despite considering simulations in excess of 200 ns (32). Additionally, lipids in the gel
phase have low rates of rotational relaxation (39). As such, on typical simulation timescales,
systems likely exist in metastable configurations that are highly dependent on the initial bilayer
structure. While this may be of little consequence for single component lipid bilayers, many
properties of multicomponent lipid mixtures depend on the in-plane morphology of the
individual lipid components, e.g., aggregation versus dispersion of lipid species in the bilayer
leaflets; furthermore, the morphology on this size scale is typically not known from experiment
(40). While self-assembled bilayers would, in principle, avoid these problems, it is not currently
practical to study the self-assembly of multicomponent bilayers with atomistically detailed
models due to the high computational cost and a rough free energy landscape that could hinder
the formation of equilibrium structures; for example, Das et al. found inverted micelle structures
formed in mixtures of 3 different CERs with CHOL and FFA unless constrained by a wall (27),
in contrast to experiments, where lamella structures form in bulk. Therefore, since preassembled
structures remain more practical for simulating atomistic models, care must be taken that system
properties are not biased by the initial configuration, so that reliable conclusions can be made.
5
Here, molecular dynamics (MD) simulations are used to examine the structural behavior
of model SC membranes, similar to those studied experimentally (19, 41), to examine the role of
the CER FA tail length and overall lipid composition on structural properties. Specifically, this
work focuses on CER NS, as this is the most abundant CER species in healthy human SC and is
typically used in model systems. Three sets of systems are studied: pure CER NS, binary
mixtures of CER NS and CHOL in 2-1 and 1-1 molar ratios, and equimolar mixtures of CER NS,
CHOL, and FFA C24:0. The structure of each of the lipids used in this work is shown in Figure
1. Mixtures of different length CERs are used in each set to examine the impact of the CER FA
tail length on the systems, which, to date, has not been the focus of any computational studies.
To ensure robust results, a random walk MD (RWMD) algorithm is validated and employed to
reduce the dependence of the final bilayer configuration on the assumed, initial morphology.
a)
HO
HO
O
b)
HO
HO
O
c)
d)
e)
NH
NH
HO
O
HO
eCER2
uCER2
CHOL
FFA
Figure 1. Chemical structure of the lipids in this study. a) CER NS C16 (eCER); b) CER NS
C24 (uCER); c) cholesterol (CHOL); d) FFA C24:0 (FFA); and e) a snapshot of a typical
configuration for the equimolar uCER2-CHOL-FFA bilayer. CER is shown in gray, CHOL is
6
shown in yellow, FFA is shown in purple, and water is shown in transparent red (oxygens) and
white (hydrogen).
MATERIALS AND METHODS
Model
CER NS consists of a sphingosine base, 18 carbons in length, linked to a saturated fatty acid
(FA) chain of variable length. Two different FA tail lengths were considered: CER NS C16:0
(Figure 1a), in which the FA tail is 16 carbons long and approximately equal in length to the
sphingosine chain (denoted as equal-length CER NS, or eCER); and CER NS C24:0 (Figure 1b),
in which the FA tail is 24 carbons long, a typical length in the SC(11) (denoted as unequal-length
CER NS, or uCER). FFA, when present in this work, is always 24 carbons in length (Figure 1d),
which has been shown to be a useful model for the FFAs in healthy human SC (11, 42). The
fully atomistic CHARMM36 force field (43) supplemented by the CHARMM-compatible CER
headgroup parameters from Guo et al. (29) were used to describe the CER NS, CHOL, and FFA.
The CER headgroup parameters have been shown to accurately reproduce the thermotropic
phase behavior of bilayers composed of pure CER NS (29), and have been used to study the
hydration of CER NS headgroups in solution (44, 45). Water was modeled using TIP3P (46).
Methods
Lipid monolayers were first constructed by placing 64 lipids on a rectangular lattice in the xy
plane with 50 Å2 per lipid. Four initial lateral distributions of a binary mixture of eCER and
CHOL (i.e., different arrangements of lipid species on the lattice) were considered for validating
the ST-based approach: completely phase-separated, consisting of one 8 x 4 block of either lipid
species (Figure 2a); a coarse-grained checkerboard, consisting of alternating 4 x 4 blocks of each
lipid species (Figure 2b); randomly mixed, where both lipid species are randomly dispersed
7
throughout the lattice (Figure 2c), and a fine-grained checkerboard, consisting of alternating lipid
species on the lattice sites (Figure 2d). To examine the impact of the CER NS FA chain length
on the structural properties, each of the four lipid compositions studied (pure CER NS, 2-1 CER
NS-CHOL, 1-1 CER NS-CHOL, and equimolar CER NS-CHOL-FFA) was subdivided into five
subcompositions, where the CER NS fraction consists of 0, 25, 50, 75, 100 mol% eCER. Thus,
in total, 20 such systems were considered. For these simulations, the effect of the initial lateral
distributions of the lipids was not explored, and the different lipid species were initially
randomly dispersed throughout the lattice, with the number of each specific lipid dictated by the
system composition.
In all cases, each lipid was rotated about its long axis by a random integer multiple of
60°, since a high degree of alignment between lipid backbones has been shown to cause
unphysically large tilt angles in the lipid tails of gel-phase bilayers (39). The monolayer was then
rotated about the x-axis, and translated in z such that the ends of the tails in opposing leaflets
were in contact, forming a bilayer; note, in this procedure each leaflet starts with the same in-
plane configuration. Twenty water molecules per lipid were then added to hydrate the outside of
the bilayers (2,560 total water molecules for 128 total lipids). System sizes ranged from 19,136
to 24,192 atoms.
The systems were relaxed through a series of energy minimization, NVT simulations, and
NPT simulations. First, a steepest descent energy minimization was performed on each initial
configuration to reduce the large non-bonded repulsions caused by (inadvertently) overlapping
atoms. Next, the systems were simulated for 10 ps in the NVT ensemble at 305 K (i.e., skin
temperature), followed by a 10 ns simulation in the NPT ensemble at 305 K and 1 atm. Unless
otherwise noted, the systems were further simulated using the RWMD algorithm. In the RWMD
8
algorithm, the system temperature is adjusted at small time intervals, such that the system takes a
random walk through temperature space. Specifically, the sequence is defined within an interval
between Tmin and Tmax, with discrete temperatures defined every !". In this work, RWMD was
performed for 50 ns, with temperature changes of !"=5×& (&∈ −1,0,1) / every 5 ps, Tmin =
305 K, and Tmax set to 355 K for the first 25 ns, and the upper bound linearly reduced to 305 K
over the final 25 ns. A representative plot of temperature versus time during the RWMD
equilibration is shown in Figure S1, where we note the sequence is defined such that it samples
all temperature states equally. Systems were further simulated at 305 K for at least 150 ns after
ST. Note that 200 ns of simulation at a fixed temperature range are used for the comparison to
the results obtained from simulations with RWMD in the RWMD validation section (i.e., for the
RWMD validation, the 305 K systems were simulated for 200 ns at 305 K, and the RWMD
systems were simulated for 200 ns with Tmin = 305 K and Tmax = 355 K). To avoid modification
to the simulation code itself, we note this random walk was determined separately and the
sequence of temperatures provided to the thermostat.
The RWMD approach is designed to mimic key aspects of the Monte Carlo simulated
tempering (ST) algorithm (47), used to find minima on a rough free energy surface by
performing a random walk through temperature space. The general idea of the ST algorithm is
that increasing the system temperature lowers free energy barriers, allowing different local
minima to be explored each time the system returns from an elevated temperature to the (lower)
temperature of interest, all while keeping the system at equilibrium due to the short timescale
(47). The latter distinguishes this approach from the widely-used simulated annealing method, in
which a system is driven out of equilibrium by increasing its temperature and simulating for a
long time at the elevated temperature, followed by a slow cooling to the temperature of interest.
9
We note that the original ST algorithm involves attempting a temperature swap after each sweep
in a Monte Carlo simulation to achieve a random walk through temperature space using the
standard Metropolis acceptance/rejection criteria, rather than the predetermined walk used in the
RWMD implementation here.
All simulations were performed in GROMACS 5.1 (48), employing the Nosé-Hoover
thermostat (49) with a coupling constant of 1 ps and, for the NPT simulations, a Parrinello-
Rahman barostat (50) with a coupling constant of 10 ps. The pressure was controlled semi-
isotropically for all NPT simulations, where the box lengths of the bilayer lateral directions were
coupled. van der Waals interactions were smoothly switched off between 10-12 Å, beyond which
they were neglected. Long-range electrostatic interactions were treated via the PME algorithm
(51) with a real-space cutoff of 12 Å. A timestep of 1 fs was used for all simulations. Note that
separate thermostats were used for the lipids and the water, although both groups follow the
same temperature walk to avoid introducing temperature gradients into the system.
Analysis
System properties were calculated over the final 100 ns of simulation at 305 K. To quantify the
structure of the bilayer, various properties were calculated, including: the area per lipid (APL),
the tilt angle of the lipid tails with respect to the bilayer normal (i.e., the z-axis), the area per tail
(APT), the nematic order parameter, density profiles of various groups across the bilayer normal,
the bilayer thickness, the width of the lipid-water interface, the width of the low density tail
region, and the hydrogen bonding. The coordination numbers (CNs) of different lipid tail pairs in
the bilayer plane are used to describe the lateral distributions of lipids; the center of mass of each
tail is projected onto the z = 0 plane for these calculations. Note that the tails were chosen for the
10
CN calculations, as the lipid tails pack in an ordered hexagonal lattice, in contrast to the
relatively more disordered packing of the lipid headgroups. The lipid backbone orientation is
used to describe the rotational motion of the lipids about their long axes. A detailed description
of each of these calculations is given in the Supporting Information.
RESULTS
Random Walk Molecular Dynamics
Before using the RWMD scheme in the rest of this work, the efficacy of the approach is first
evaluated for equimolar mixtures of eCER and CHOL. This composition was chosen as eCER
and CHOL have similar hydrophobic lengths, which should yield bilayers without a considerable
interdigitation region in the middle; prior work has shown that CHOL can reside in this
interdigitation region in lipid mixtures with uCER (28). Thus, the lipids (especially CHOL) in
this mixture should stay in their canonical bilayer conformation, i.e., with their headgroups at the
lipid-water interface and their tails creating a hydrophobic core, therefore allowing evaluation of
the ability of RWMD to enhance the in-plane lateral and rotational rearrangements of the lipids.
Figure 2 shows the initial configuration of each system studied, the final configurations at 200
ns at 305 K, the final configurations after 200 ns of RWMD, and the evolution of the CHOL-
CHOL CNs for each equilibration scheme. Note that the CHOL-CHOL CN was chosen because
it is the least biased CN (e.g., the FA-SPH CN is skewed because the FA and SPH tails are part
of the same molecule, and hence will show a high level of association). Since lipids of the
equimolar eCER-CHOL mixture are expected to mix, the fully separated and coarse-grained
checkerboard morphologies (Figures 2a and 2b) serve as "bad" initial configurations, and the
randomly mixed and fine-grained checkerboard systems (Figures 2c and 2d) serve as naïve, but
11
reasonable, guesses of how to initialize mixed-lipid bilayers. Each initial morphology was
equilibrated for 200 ns with two different schemes: standard MD at 305 K, and RWMD with 305
K < T < 355 K. Both visual inspection and the evolution of the CHOL-CHOL CN can be used to
compare the in-plane morphologies, and hence equilibration, of the different systems and
equilibration procedures. At 305 K, the final morphologies visually resemble the corresponding
initial morphologies; this is reflected in the steady-state nature of the CNs at 305 K, which are
listed in Table 1. This is especially notable for the fully-separated and coarse-grained
checkerboard systems, where the large CHOL aggregates make it easy to visually establish that
there are only slight changes in the shape of the initial aggregates. These results highlight the
frozen nature of these systems at 305 K; only small rearrangements occur in 200 ns if RWMD is
not applied. Note that in the fine-grained (FG) checkerboard system, small linear aggregates of
CHOL form relatively quickly, indicated by the increase in the CN during the first 25 ns,
although the system does not evolve much after this initial change.
12
Figure 2. System configurations and CHOL-CHOL coordination number (CN) as a function of
initial configuration and equilibration procedure for the equimolar mixture of eCER2 and CHOL.
The first column shows the different initial configurations: a) maximum phase separation; b)
coarse-grained checkerboard; c) randomly mixed; and d) fine-grained checkerboard. The second
column shows the systems after 200 ns of MD at 305 K, the third column shows the systems
after 200 ns of RWMD, and the fourth column shows the evolution of the CHOL-CHOL CN for
each initial configuration. Snapshots are taken along the z-axis, i.e., along the bilayer normal
direction. eCER2 is represented as blue spheres, CHOL is represented as yellow spheres. For
clarity, water is not shown.
Table 1. CHOL-CHOL coordination number over the final 50 ns of MD simulation. Separated
refers to the initial configuration shown in Figure 2a, CG Checker refers to that in Figure 2b,
Random Mix refers to Figure 2c, and FG Checker refers to Figure 2d. The uncertainties here, and
in all subsequent tables and figures, are given as the standard error of the measured property.
13
Initial Configuration
Equilibration Separated CG Checker Random Mix FG Checker
305 K
1.75 ± 0.07
2.5 ± 0.1
RWMD
4.46 ± 0.03 3.72 ± 0.04
2.3 ± 0.1
2.4 ± 0.1
2.53 ± 0.08
2.4 ± 0.2
With RWMD, the final lateral distributions of the lipids are visually distinct from the initial
configuration, showing that the lipids were able to reorganize within the bilayer leaflets. For the
two most separated systems, the larger CHOL aggregates mostly break up into smaller
aggregates (Figure 2a, i and Figure 2b, j), while small CHOL aggregates form in the fine-
grained checkerboard system (Figure 2d, l). Importantly, all of the final configurations after
RWMD visually resemble each other and have nearly identical CN values, as listed in Table 1,
and systems that start close to the final morphology (i.e., the random and fine-grained
checkerboard) rapidly converge to their final coordination number. Since each of these systems
started from qualitatively different initial configurations, this result shows that the final
configurations are decorrelated from the initial configurations and RWMD can reproducibly
form the structures with matching properties without strong bias from the starting configuration.
We note this does not necessarily indicate that RWMD has found the true free-energy minimum
of the system, but this does strongly suggest that it is a stable, low energy configuration, given
that multiple independent trajectories converge to the same state.
Analysis of the rotational motion of the eCER molecules also suggests that the systems
equilibrated with RWMD are more decorrelated from their initial configurations than those
equilibrated at 305 K. Figure 3 shows the autocorrelation function of the lipid backbone angle in
the bilayer plane for the different initial configurations and equilibration methods. The lipid
14
backbone orientations become completely uncorrelated from the initial orientations by 25 ns of
RWMD, while slower relaxation is observed for the systems simulated at 305 K. Interestingly,
the phase-separated system shows the slowest relaxation of the eCER backbone orientations,
likely because the large, dense eCER domain has a larger energy barrier for lipid rotation due to
lipid-lipid hydrogen bonding. Since the systems that were equilibrated at 305 K all converge to
different morphologies and the lipid backbone orientations are correlated to the initial
orientations, at least three of the four systems must be metastable configurations. Often in bilayer
simulations, a specific property, such as the area per lipid (APL), is monitored over time, and
equilibrium is assumed when that property reaches a steady state. This practice, however, may
lead to spurious assumptions of equilibrium since other features of the bilayer structure may not
exist in their preferred state, but this cannot be easily verified since the preferred state is not
always known a priori. For example, the APLs of the systems equilibrated at 305 K are
compared to the APLs of the systems equilibrated with RWMD (after running at 305 K for a
more direct comparison) in Table 2, from which it is apparent that the systems equilibrated at
305 K show a larger spread in the APLs than the systems equilibrated with RWMD. The two
systems with the highest level of CHOL aggregation have similar APLs, while the other two
systems have lower APLs that are similar to each other. Due to the different morphologies and
packing densities of the lipids in these systems, they likely have different properties, e.g.,
permeability, which is related to the in-plane density of the lipids (52). In contrast, all of the
systems equilibrated with RWMD have APLs consistent within the standard deviation, further
illustrating the reproducibility, and hence confidence gained, when equilibrating with RWMD.
Table 2. Area per lipid (in Å2) of systems as a function of initial configuration and equilibration
methodology. The systems equilibrated with RWMD were simulated at 305 K for 50 ns for a
15
more direct comparison with the other systems. Separated refers to Figure 2a, CG Checker refers
to Figure 2b, Random Mix refers to Figure 2c, and FG Checker refers to Figure 2d.
Initial Configuration
Equilibration Separated CG Checker Random Mix FG Checker
305 K
RWMD
38.8 ± 0.4 38.8 ± 0.4
39.6 ± 0.5 39.7 ± 0.4
39.4 ± 0.2
39.5 ± 0.4
39.6 ± 0.2
39.4 ± 0.3
Figure 3. Autocorrelation function C(t) of the eCER backbone orientations during simulations at
305 K and with RWMD from different initial configurations. Separated refers to Figure 2a, CG
Checker refers to Figure 2b, Random Mix refers to Figure 2c, and FG Checker refers to Figure
2d. The shaded area shown is the standard deviation over the rotational autocorrelation function
of all eCER molecules in the bilayer.
16
This lack of mobility and difficulty in proper equilibration is a recognized problem when
simulating CER-based bilayers, with many studies running the systems at elevated temperatures
in an attempt to overcome mobility issues (25-28, 30, 31, 34, 35). Here, we also compare to
equilibration and production runs at 340 K, as several MD studies of SC lipids have used this
temperature (25-28). Figure 4 shows the evolution of the CN for the fully separated system (Fig.
2a) simulated for 600 ns at 305 K, 340 K, and with RWMD. As expected, the system at 305 K is
still frozen in simulations in excess of 0.5 µs, while the system simulated with RWMD reaches
its steady state CN number after ~110 ns. Perhaps unsurprisingly, the system simulated at 340 K
shows behavior somewhere in between that of the other systems; while lateral reorganization
does occur, the CN reaches a steady state value after ~520 ns, which we note is significantly
longer than the simulation times typically studied. Note that this system also reaches a pseudo-
steady state CN between 80 and 150 ns; with many studies performed on simulations run for 100
ns or less, this metastable state could be erroneously mistaken for equilibrium, if one were
relying on CN to determine convergence. At 600 ns, the rotational relaxation of the lipids for this
system at 340 K is higher than 305 K, but still lower than RWMD (Figure 3a). While the fully
phase-separated system is likely a poorly chosen starting configuration given some level of prior
knowledge of this particular system, we emphasize that the actual morphology is typically not
known a priori. Thus, these results indicate that RWMD is more efficient than simply running at
a high temperature to equilibrate multicomponent, gel-phase lipid bilayers and can provide
increased confidence in the reproducibility of the results, particularly for systems in which the
lateral organization is unknown and expected to play an important role in properties of interest.
17
Figure 4. Evolution of the CHOL-CHOL coordination number (CN) for the fully phase-
separated systems equilibrated with three different methodologies as a function of time (t).
Dependence of Structural Properties on Bilayer Composition
The effect of the composition of CER-based lipid bilayers on structural properties is now
examined. Note, all systems start from a randomly dispersed morphology, given that, in the prior
section, this arrangement relaxed the fastest and appears to be reasonably representative of the
morphology for this family of systems; these systems were also relaxed using RWMD as
described in Methods.
General Structural Properties. All systems exist in dense, highly ordered, gel phases, as
indicated by high nematic order parameters ranging from 0.95 to 0.99, shown in Figure S4 and
Table S1. The addition of CHOL decreases the nematic order, consistent with Das et al. (26), but
the magnitude of this decrease is quite small, (e.g., ranging from 0.988 for pure uCER to 0.954
for 1-1 CER-CHOL with 3-1 eCER-uCER) and the lipids are still highly ordered from visual
inspection (Figure 1e), indicating no phase change upon addition of CHOL for these
compositions.
18
The tilt angle shows a weak dependence on composition, as shown in Figure 5a. The tilt
angle depends significantly on the CER FA tail length for the pure CER systems, where a strong
increase is seen with increasing eCER for systems with 50% or more eCER. This trend is a result
of the balance between headgroup and tail interactions: steric repulsions between lipid
headgroups dictate the APL, while van der Waals attractions between the lipid tails cause them to
tilt to optimize their spacing (53). For the CER bilayers that are predominantly uCER, the tail
length asymmetry gives rise to a wide low density tail region in the center of the bilayer
(discussed below), which increases the effective optimal tail packing, leading to a lower tilt
angle. The addition of CHOL and FFA causes the tails to tilt less and show almost no change as
a function of increasing eCER, likely because CHOL has a bulky ring structure with a small
headgroup, and therefore acts as a spacer between the lipid tails, buffering the head-tail effects
seen in the pure systems. Hence, there is a smaller mismatch between the packing densities of the
headgroups and tails, leading to a smaller tilt. Notably, the 1-1 CER-CHOL and ternary mixtures
have very similar tilt angles, which we attribute to the fact that these systems have a similar
fraction of alkyl tails.
19
Figure 5. (a) Average tilt angle (q) of the lipid tails with respect to the bilayer normal, (b) area
per lipid (APL) and (c) area per tail (APT) as a function of eCER2 composition. Note that the
abscissa is the fraction of CER NS that is eCER, not the total eCER fraction in the system.
The APL and APT trends as a function of the fraction of eCER are shown in Figure 5b,c.
While the APL describes the density of lipids in the bilayer plane, irrespective of the number of
tails that each lipid has, the APT describes the tail packing density in the plane of the lipid tails
(i.e., APT accounts for both the number of tails and the tilt angle). All CER NS and CER NS-
CHOL systems have similar APLs, between 38-40 Å2, as expected since CER and CHOL have
similar cross-sectional areas (54). There is only a slight, nonmonotonic dependence of the APL
on the CER FA tail length. This trend is expected since the headgroups dictate the APL and these
systems have the same headgroups for a given composition. Systems containing FFA have
smaller APLs, since FFA contains a single alkyl chain. The APT increases with the CHOL
content, as CHOL and CER NS have similar cross-sectional areas, but CHOL is treated as a
20
(DMPC, a phospholipid with
single tail. The ternary systems have APTs more similar to the 2-1 CER NS-CHOL systems,
which is reasonable considering they have the same ratio of CHOL to alkyl chains. The mixed
lipid systems all show a slightly increasing APT with increasing eCER fraction. Interestingly,
there appears to be some level of APT non-additivity in these systems. For example, consider the
APL and APTs of the 1-1 CER NS-CHOL systems. Since they have similar APLs as the pure
CER NS systems, but ¾ the number of tails, ideal mixing would yield APL values that are 4/3
that of the pure CER NS values. However, the actual values are lower, suggesting a relative
attraction between CHOL and the saturated tails of CER NS, similar to what was observed for
dimyristoylphosphocholine
tails), but not
dioleoylphosphocholine (DOPC, a phospholipid with unsaturated tails) (55). Although perhaps
expected based on chain saturation, this behavior is nontrivial for CER systems, since the CER
NS headgroups are much smaller than PC headgroups and CHOL presumably has less empty
space to occupy.
Density Profiles. The bilayer thickness, calculated from the water density profiles, is a function
of both overall lipid composition and CER FA tail length (Figure S5 and Table S2). The bilayer
thickness decreases with increasing eCER and CHOL content, as one would expect since these
lipids have shorter tail lengths than uCER or FFA. However, the tail length asymmetry leads to a
richer behavior, which can be observed via examination of the lipid density profiles, as shown
for each system in Figure 6.
saturated
The pure CER bilayers exhibit mass density profiles that contain the expected features of
a lipid bilayer. Near the lipid-water interface, the profiles contain a peak (Figure 6a), which
represents the heavy atoms in the CER headgroups. Just inside the headgroup region, there is a
high density tail region where the tails are densely packed and highly ordered (see discussion of
21
nematic order above), consistent with prior work (37). In the center of the bilayer is a low
density tail region, where the lipid tails are less densely packed and less ordered. This region
exists primarily due to the asymmetry in the lengths of the tails in the systems and is composed
of the terminal part of the uCER tail and FFA tails, as shown in Figure S6. In the systems with
less asymmetry (i.e., high eCER and CHOL concentrations), the profile is more V-shaped in the
middle since there is little interdigitation between tails in opposing leaflets. Comparing the high
and low density tail regions for the different bilayers, the width of the high density tail region is
found to be nearly constant across all compositions, whereas the width of the low density tail
region varies with eCER and CHOL content. Figure 7 shows the total bilayer thickness as a
function of the thickness of this low density tail region in the center of the bilayer. Clearly, the
thickness of the low density tail region dictates the total bilayer thickness for a given
composition. The width of the high density tail region is determined by the smallest hydrophobic
length in the system, where the tails must pack densely to fit in the area dictated by the
headgroups. For the lipids in this study, this is roughly 16 carbons, or the length of the
sphingosine chain that is not part of the CER NS headgroup, which is also similar to the length
of a CHOL molecule (56).
22
The profiles for the 5 subcompositions are shown for all compositions and are labelled as
Figure 6. Total lipid mass density profiles across the bilayer for all systems considered: a) pure
CER NS, b) 2-1 CER NS-CHOL, c) 1-1 CER NS-CHOL, d) equimolar CER NS-CHOL-FFA.
follows: Solid black line 100 mol% eCER, dashed red line 75 mol% eCER, dot-dashed blue line
50 mol% eCER, solid orange line 25 mol% eCER, dotted purple line 0 mol% eCER. Note that
these percentages denote the fraction of CER NS that is eCER, not the total fraction of eCER in
the system.
23
Figure 7. Total bilayer thickness as a function of the thickness of the low density tail region,
with the different points on each line denoting the different eCER fractions; the smallest values
indicate 100 mol% eCER, and the largest 0 mol% eCER.
The presence of CHOL and FFA has two main effects on the density profiles. First, the
smaller headgroups of CHOL and FFA lead to smaller peaks in the headgroup region, compared
to a pure CER NS bilayer; these peaks are dramatically reduced going from pure CER to 2-1
CER NS-CHOL and are mostly absent in the 1-1 CER NS-CHOL and ternary systems. Second,
the width of the low density tail region increases with the addition of CHOL and FFA, as shown
in Figure 6 and Figure 7. Additionally, the low density tail regions tend to have more features in
the mixed lipid systems, e.g., shoulders representing intermediate densities due to the different
tail lengths. Peaks also appear in the middle of the density profiles for the mixed systems, due
both to the long FA chain of uCER (and FFA for the 3 component system) and the presence of
CHOL that has migrated from the ordered bilayer region into the low density tail region (28). In
the eCER-CHOL systems, the lack of significant interdigitation means this peak is absent;
24
however, all systems containing FFA demonstrate a peak in the middle of the density profile
because of the long tail of FFA.
Figure 8 shows the density profiles for the ternary system with 1-1 eCER-uCER, along
the bilayer normal direction, broken down by the contribution of specific lipid components. The
CHOL headgroup sits ~3 Å deeper into the bilayer with respect to the CER headgroups, which is
consistent with prior simulation and experimental studies (23, 26, 31). There is also a small
subpopulation of CHOL lying flat in the middle of the bilayer, illustrated by the peak in the
middle of the CHOL ring density profile, which is consistent with the work of Das et al. (28)
Additionally, the FFA headgroups tend to sit slightly further into the water than the CER
headgroups.
Figure 8. Mass density profiles of various groups in the ternary bilayer with a 1-1 eCER2-
uCER2 composition. Note that the profiles for different groups are shifted vertically for clarity.
Each curve is offset by 0.1 g/mL for each subsequent dataset in the legend.
The interfacial width, which can indicate the hydrophobicity of the surface (e.g., a larger
interfacial width would indicate a more hydrophilic surface), is defined as the width of the region
over which the water density drops from the bulk value to 1/e of the bulk value. The bilayers in
this work appear hydrophobic, having interfacial thicknesses between 3.5-6 Å2, compared to
25
phospholipid bilayers, which are generally greater than 10 Å (57). There is little dependence on
the bilayer composition or CER FA tail length, as shown in Figure S7 and Table S3. Thus, we
conclude that while the lamellar organization near the middle of the bilayer depends on the lipid
composition, the apparent hydrophobicity of the lipid-water interface does not.
Hydrogen Bonding. The total number of lipid-lipid hydrogen bonds in each system is
listed in Table 3. The lipid-lipid hydrogen bonding is a function of the total CER content, with a
reduction in CERs resulting in fewer lipid-lipid hydrogen bonds because CHOL and FFA
molecules have fewer hydrogen bonding sites. The lipid-lipid hydrogen bonding is also
independent of the eCER fraction, which may be expected since eCER and uCER have identical
headgroups and their relative position at the interface is essentially unchanged as a function of
the eCER2 fraction.
26
Table 3. Total number of lipid-lipid and lipid-water hydrogen bonds for each system studied.
i
d
p
i
l
-
d
i
i
p
L
r
e
t
a
w
-
d
p
L
i
i
Fraction eCER
0.25 0.50 0.75 1.0
0
Composition
62
63
63
Pure CER NS
38
37
2-1 CER NS-CHOL 40
26
27
1-1 CER NS-CHOL 28
19
19
19
Ternary
173 168 173 176 171
Pure CER NS
2-1 CER NS-CHOL 173 169 173 175 176
1-1 CER NS-CHOL 168 170 164 167 173
154 152 158 158 155
Ternary
61
38
28
19
63
39
28
19
The hydrogen bonding between specific pairs of lipids for the ternary mixture with 1-1
eCER-uCER is listed in Table S4. Negligible CHOL-CHOL hydrogen bonding is observed,
which can be rationalized by the fact that CHOL only has 1 hydrogen bond donor and acceptor,
and two neighboring CHOL molecules would have to adopt a strained configuration to form a
hydrogen bond. There are also very few CHOL-FFA hydrogen bonds, likely a result of CHOL
sitting deeper into the bilayer. An appreciable number of CER-CHOL hydrogen bonds are seen,
with slightly more eCER-CHOL (4.3) than uCER-CHOL (3.7) hydrogen bonds. Interestingly,
the level of CER-CHOL hydrogen bonding is similar in the 2-1 and 1-1 CER-CHOL systems
with, for example, 11.4 and 11.1 total CER-CHOL hydrogen bonds for the 2-1 and 1-1 CER-
CHOL systems with 50 mol% eCER (Table S5 and Table S6); this trend may suggest that the
CER-CHOL hydrogen bonding is saturated at or below 33 mol% CHOL. Additionally, when
27
CER is replaced by CHOL (i.e., comparing the pure CER NS and the 2-1 CER NS-CHOL or 1-1
CER NS-CHOL systems), there are fewer hydrogen bonds relative to the total number of
hydrogen bonding sites in the system, indicating that CER-CER hydrogen bonds are preferred
over CER-CHOL hydrogen bonds. There are comparatively fewer CER-CHOL hydrogen bonds
in the ternary systems, since FFA is present and is competing with CHOL to form hydrogen
bonds with the CER. There is a surprisingly low amount of FFA-FFA hydrogen bonding, given
that FFA prefers to be near other FFA molecules (discussed below); however, FFA sits deeper
into the water and thus forms more hydrogen bonds with water.
The amount of lipid-water hydrogen bonding in each system is also listed in Table 3.
Despite the relatively hydrophobic lipid-water interface, there are significantly more lipid-water
hydrogen bonds than lipid-lipid hydrogen bonds. As with the lipid-lipid hydrogen bonds, the
lipid-water hydrogen bonding is independent of the CER FA tail length. Interestingly, the pure
CER and 2-1 CER-CHOL systems have similar amounts of lipid-water hydrogen bonds, despite
the fact that CHOL has significantly fewer hydrogen bonding sites than CER, while the 1-1
CER-CHOL and ternary systems have comparably less lipid-water hydrogen bonds, since these
systems have less CER (and hence fewer hydrogen bonding sites) than the systems with other
two compositions. This trend in lipid-water hydrogen bonding is perhaps explained by the
"spacer" effect of CHOL, discussed above with respect to the tilt angles. Although less CER
results in fewer available hydrogen bonding sites, the spacing effect of CHOL allows the water
to more thoroughly hydrate the CER headgroups and hence results in more lipid-water hydrogen
bonds per hydrogen bonding site.
In-Plane Morphology. For a general view of the in-plane morphology of each system, we
examine coordination numbers (CNs) between specific lipid tails in the bilayer plane. The
28
CHOL-CHOL CNs depend on the amount of CHOL and available neighbors in a given system,
and not on the CER FA tail length, as detailed in Figure S8 and Table S7. This result suggests a
random distribution of CHOL throughout the bilayer leaflets. For example, there are equal
numbers of three different types of tails in the 1-1 CER-CHOL systems (i.e., CER-FA, CER-
SPH, and CHOL). Random mixing would give the observed CHOL-CHOL CN of 2. Applying
this logic to the 2-1 CER-CHOL and ternary systems, random mixing would imply CHOL-
CHOL CNs of 1.2 and 1.5 for each composition, respectively, which is observed. In contrast to
CHOL, FFA shows a preference for specific neighbors, and thus has a less random distribution
throughout the bilayer leaflets. In the uCER-CHOL-FFA system, the FFA-CERFA and FFA-
CERSPH CNs are 1.39 ± 0.08 and 1.20 ± 0.07, respectively (where CERFA and CERSPH represent
the fatty acid and sphingosine tails of CER, respectively). This result suggests that FFA has a
preference for the FA chain of CER compared to the SPH chain. Additionally, a significant
preference for FFA to be near FFA is observed, with a FFA-FFA CN of 2.3 ± 0.1, compared to
1.39 ± 0.08 for FFA-CERFA, the second highest. This has also been observed experimentally,
with FFA-enriched domains forming with the tails tightly packed on an orthorhombic lattice
(19). However, unlike experimental results on similar systems (41), we do not see any chain
length-dependent behavior in regards to the mixing of CER and FFA. We note though that
comparable experimental systems consider multilamellar structures, whereas we consider a
single hydrated bilayer, which could account for the differences observed. We also note that the
systems studied in the current work are several orders of magnitude smaller than comparable
experimental systems (nm vs. µm), and that an in-depth study of lateral distributions of lipids is
beyond the scope of this work.
29
DISCUSSION
Relevance to Experimental Models of the Short Periodicity Phase. The localization of different
groups within an experimental model of the short periodicity phase (SPP) of the SC has been
studied with neutron diffraction (22, 23). These experimental systems resemble the systems
studied here, with a few differences. First, we only consider CER NS, whereas experimental
systems tend to include a mixture of CERs with different headgroups, although uCER typically
accounts for the majority (~60%) of the CERs (11, 23). Secondly, we also only consider FFA
C24, whereas experimental systems often include FFAs with a distribution of tail lengths. We
note, however, recent work has shown that CER headgroup chemistry does not play a strong role
in the lamellar organization of SPP models, and that systems with a distribution of FFA tail
lengths had the same lamellar organization as a similar system with only FFA C24 (11). Perhaps
the biggest difference between experimental systems and the systems studied here is that we
consider a single bilayer, which is the typical approach to studying these systems with molecular
simulation. Experimental systems usually contain membranes with numerous diffraction orders
detected, indicating multilayer structures. Nonetheless, we can directly compare neutron
scattering length density (NSLD) profiles from simulation and experiment. Additionally, we can
compare the localization of specific groups based on mass density profiles from simulation and
those reconstructed from selective deuteration of specific groups from experiment.
In Figure 9 we compare the total NSLD profiles from experiment (22, 23) and
simulation. The most obvious features of each are the large peaks at ±27 Å (experiment) and ±25
Å (simulation). Given the experimental resolution of ~5.4 Å, the locations of these peaks are in
good agreement. Both simulation and experiment show valleys in the NSLD at ~7 Å, which we
note aligns with the dips in the low tail density region from Figure 6d. Additionally, subtle
30
shoulders in the NSLDs are present at ±15 Å in both simulation and experiment, which aligns
with the edges of the low tail density region shown in Figure 6d.
Figure 9. Neutron scattering length density (NSLD) profiles, reconstructed from experiment(22)
and calculated from simulation. Note that the profiles were shifted such that the minimum lies at
0, and the simulation curve was scaled to have the same height as the experimental curve.
Focusing next on the localization of specific groups within the bilayer, we compare the
mass density profiles from simulation with the difference between the protonated and deuterated
NSLDs from experiment, shown in Figure S9. In this manner, we can compare the localization of
specific groups between simulation and experiment. Comparing the location of the FA tail of
uCER that was measured in Groen et al., (22) we find good agreement, with a single broad peak
spanning ±7.5 Å (Figure S9a), which corresponds to the region where the uCER FA tails from
opposing leaflets interdigitate. The CHOL tails are localized at ±2 Å in both simulation and
experiment (Figure S9b) (23). The CHOL headgroups show some deviation between experiment
and simulation, localized at ±23 Å and ±21 Å, respectively (Figure S9c). This discrepancy,
however, is again within the experimental resolution of 5.4 Å (23). Despite the fact that we are
31
studying single bilayers, the locations of specific groups within the bilayer agree very well with
experimental data on systems with similar compositions. Therefore, we can conclude that the
model systems accurately approximate the model SPP systems from experiment, and conclusions
from this work also likely apply to model SPP systems, and also the SPP in the SC.
Effects on Barrier Properties of SC. In this work, the lipid composition and CER FA tail
length were found to have the largest impact on lamellar organization, with the bilayer thickness
and shape of the density profile in the low density tail region most affected. If we consider how
these changes would affect the barrier properties (i.e., the permeability) of these bilayers, we can
provide some insight into the observed experimental behavior. The permeability of a membrane
is a product of the solute diffusion and partitioning, where the partitioning has an exponential
dependence on the free energy of solvation. Due to this exponential dependence, small changes
in local packing (i.e., depth-dependent density), can have a dominant impact on the bilayer
permeability. Since the lipid mass density profiles are not constant across the bilayers, the
resistance experienced by a permeant molecule would vary with bilayer depth. Moreover, since
the low density tail regions are qualitatively different for the bilayers with different
compositions, these depth-dependent partition coefficients would also change, albeit nontrivially.
Thus, if one was interested in comparing the permeability of these different systems, a simple
homogeneous solubility-diffusion model would not suffice. This indeed is not unrecognized in
the field, as the few simulation studies of SC lipid permeability in the literature to date have used
an inhomogeneous solubility-diffusion model (30, 52).
While the structure of the lipids at the lipid-water interface changes with composition, it
does not change with CER FA tail length. This is illustrated by the fact that the thickness of the
interfacial region and lipid density profiles near the headgroups are unchanged with the CER FA
32
tail length for a given composition (Figure 6 and Figure S7). Additionally, both the total number
of lipid-water and lipid-lipid hydrogen bonds is constant for a given composition. Thus, we
expect any changes in the barrier properties of model SC membranes with CER FA tail length to
be a result of changes in the structure of the bilayers in the hydrophobic core, and that changes in
the headgroup region would play only a small role. This observation is consistent with the
permeability measurements reported by Uchiyama et al., which show that the permeability of
ethyl-p-aminobenzoic acid through membranes composed of synthetic mixtures of CER, CHOL,
and FFA is independent of the CER composition, but strongly dependent on the FFA tail length
dispersity (11).
CONCLUSIONS
We have studied via molecular dynamics simulations the structure of lipid bilayers relevant to
the stratum corneum layer of the skin, as well as proposed and validated a thorough relaxation
protocol for such systems. We first illustrated several difficulties in simulating SC lipid mixtures
and multicomponent gel-phase bilayers in general e.g., systems tend to be frozen at physiological
temperatures with only small lateral rearrangements possible. Since the most realistic
morphologies are not generally known a priori, systems must be allowed to find the most
realistic morphology. This task is shown to be computationally impractical at physiological
temperatures and inefficient at commonly used elevated temperatures, as well as likely to lead to
metastable states being confused for equilibrium states. The proposed random walk MD
methodology, based on simulated tempering, was shown to be an efficient and reproducible
method for minimizing the influence of an assumed, initial membrane configuration for
multicomponent, gel-phase bilayers. We therefore expect this method to be especially useful as
33
simulations incorporate more complex lipid mixtures. Using the random walk MD protocol, we
examined a series of SC lipid bilayers with varying lipid compositions and CER FA tail length.
We showed that the lamellar organization is most affected by the lipid composition and CER FA
tail length. Subtle changes to the lateral organization of the lipid tails as a function of the CER
FA tail length were also seen. Additionally, since it was observed that the behavior at the lipid-
water interface does not change with the CER FA tail length, we speculate that any tail length-
dependent changes in barrier function are a result of changes in the lipid tail region and not the
headgroup region.
34
Supporting Information. The Supporting Information contains the content listed below.
An explanation of the various metrics used to quantify the structure of the bilayers in this study.
Figure S1.
Representative plot of temperature versus time during RWMD equilibration.
Figure S2.
Schematic of the bilayer thickness and interfacial thickness calculations.
Figure S3. Visualization of the low tail density region.
Figure S4. Nematic order parameter of lipid tails.
Figure S5.
Bilayer thickness as a function of lipid composition.
Figure S6. Mass density profiles of the "unequal" part of the uCER2 tail and terminal 8
carbons of the FFA tail in the equimolar uCER2-CHOL-FFA system.
Figure S7.
Thickness of lipid water interface as a function of system composition.
Figure S8. CHOL-CHOL coordination number as a function of system composition.
Figure S9.
Localization of various lipid components compared to neutron scattering results.
Table S1.
Nematic order parameter of lipid tails as a function of bilayer composition.
Table S2.
Bilayer thickness as a function of bilayer composition.
Table S3.
Thickness of lipid-water interface as a function of bilayer composition.
Hydrogen bonding, broken down by lipid component, for the ternary system with
Table S4.
a 1-1 eCER2-uCER2 ratio.
35
Hydrogen bonding, broken down by lipid component, for the 2-1 CER2-CHOL
Table S5.
system with a 1-1 eCER2-uCER2 ratio.
Hydrogen bonding, broken down by lipid component, for the 1-1 CER2-CHOL
Table S6.
system with a 1-1 eCER2-uCER2 ratio.
Table S7.
CHOL-CHOL coordination numbers as a function of system composition.
AUTHOR CONTRIBUTIONS
C.M., C.R.I., and A.L.B. designed research; T.C.M., and C.R.I. performed research; T.C.M. and
R.H. contributed analytic tools; all authors analyzed data and wrote the article.
ACKNOWLEDGEMENTS
This work was supported by grant number R01 AR057886-01 from the National Institute of
Arthritis and Musculoskeletal and Skin Diseases and National Science Foundation grant number
CBET-1028374. This work was conducted in part using computational resources provided by the
National Energy Research Scientific Computing Center, supported by the Office of Science of
the Department of Energy under Contract No. DE-AC02-05CH11231 and the Advanced
Computing Center for Research and Education at Vanderbilt University.
REFERENCES
1.
2.
Madison, K. C. 2003. Barrier Function of the Skin: ''La Raison d'Etre'' of the
Epidermis. The Journal of Investigative Dermatology 121:231-241.
Weerheim, A., and M. Ponec. 2001. Determination of stratum corneum lipid profile by
tape stripping in combination with high-performance thin-layer chromatography.
Archives of Dermatological Research 293:191-199.
36
3.
4.
5.
6.
7.
8.
Janssens, M., J. van Smeden, G. S. Gooris, W. Bras, G. Portale, P. J. Caspers, R. J.
Vreeken, T. Hankemeier, S. Kezic, R. Wolterbeek, A. P. M. Lavrijsen, and J. A.
Bouwstra. 2012. Increase in short-chain ceramides correlates with an altered lipid
organization and decreased barrier function in atopic eczema patients. The Journal of
Lipid Research 53:2755-2766.
Lavrijsen, A. P. M., J. A. Bouwstra, G. S. Gooris, A. Weerheim, H. E. Bodde, and M.
Ponec. 1995. Reduced Skin Barrier Function Parallels Abnormal Stratum Corneum Lipid
Organization in Patients with Lamellar Ichthyosis. The Journal of Investigative
Dermatology 105:619-624.
Pilgram, G. S. K., D. C. Vissers, H. van der Meulen, S. Pavel, A. P. M. Lavrijsen, J. A.
Bouwstra, and H. K. Koerten. 2001. Aberrant lipid organization in stratum corneum of
patients with atopic dermatitis and lamellar ichthyosis. The Journal of Investigative
Dermatology 117:710-717.
van Smeden, J., M. Janssens, W. A. Boiten, V. van Drongelen, L. Furio, R. J. Vreeken,
A. Hovnanian, and J. A. Bouwstra. 2014. Intercellular skin barrier lipid composition and
organization in Netherton syndrome patients. The Journal of Investigative Dermatology
134:1238-1245.
Ishikawa, J., H. Narita, N. Kondo, M. Hotta, Y. Takagi, Y. Masukawa, T. Kitahara, Y.
Takema, S. Koyano, S. Yamazaki, and A. Hatamochi. 2010. Changes in the ceramide
profile of atopic dermatitis patients. The Journal of Investigative Dermatology 130:2511-
2514.
Joo, K.-M., J.-H. Hwang, S. Bae, D.-H. Nahm, H.-S. Park, Y.-M. Ye, and K.-M. Lim.
2015. Relationship of ceramide-, and free fatty acid-cholesterol ratios in the stratum
corneum with skin barrier function of normal, atopic dermatitis lesional and non-lesional
skins. Journal of Dermatological Science 77:71-81.
Rabionet, M., K. Gorgas, and R. Sandhoff. 2014. Ceramide synthesis in the epidermis.
Biochimica et Biophysica Acta - Molecular and Cell Biology of Lipids 1841:422-434.
de Jager, M., W. Groenink, R. Bielsa i Guivernau, E. Andersson, N. Angelova, M. Ponec,
and J. A. Bouwstra. 2006. A novel in vitro percutaneous penetration model: evaluation of
barrier properties with p-aminobenzoic acid and two of its derivatives. Pharmaceutical
Research 23:951-960.
Uchiyama, M., M. Oguri, E. H. Mojumdar, G. S. Gooris, and J. A. Bouwstra. 2016. Free
fatty acids chain length distribution affects the permeability of skin lipid model
membranes. Biochimica et Biophysica Acta (BBA) - Biomembranes 1858:2050-2059.
Garidel, P., B. Fölting, I. Schaller, and A. Kerth. 2010. The microstructure of the stratum
corneum lipid barrier: mid-infrared spectroscopic studies of hydrated ceramide:palmitic
acid:cholesterol model systems. Biophysical Chemistry 150:144-156.
13.
Lafleur, M. 1998. Phase behaviour of model stratum corneum lipid mixtures: an infrared
spectroscopy investigation. Canadian Journal of Chemistry 76:1501-1511.
14. Mojumdar, E. H., G. S. Gooris, and J. A. Bouwstra. 2015. Phase behavior of skin lipid
mixtures: the effect of cholesterol on lipid organization. Soft Matter 11:4326-4336.
15. Mojumdar, E. H., G. S. Gooris, D. Groen, David J. Barlow, M. J. Lawrence, B. Demé,
and J. A. Bouwstra. 2016. Stratum corneum lipid matrix: location of acyl ceramide and
cholesterol in the unit cell of the long periodicity phase. Biochimica et Biophysica Acta
(BBA) - Biomembranes 1858:1926-1934.
9.
10.
11.
12.
37
21.
22.
19.
20.
16. Moore, D. J., M. E. Rerek, and R. Mendelsohn. 1999. Role of ceramides 2 and 5 in the
structure of the stratum corneum lipid barrier. International Journal of Cosmetic Science
21:353-368.
17. Mueller, J., A. Schroeter, R. Steitz, M. Trapp, and R. H. H. Neubert. 2016. Preparation of
a new oligolamellar stratum corneum lipid model. Langmuir 32:4673-4680.
Rowat, A. C., N. Kitson, and J. L. Thewalt. 2006. Interactions of oleic acid and model
18.
stratum corneum membranes as seen by 2H NMR. International Journal of Pharmaceutics
307:225-231.
Školová, B., B. Januìššová, J. Zbytovská, G. S. Gooris, J. A. Bouwstra, P. Slepička, P.
Berka, J. Roh, K. Palát, A. Hrabálek, and K. Vávrová. 2013. Ceramides in the skin lipid
membranes: Length matters. Langmuir 29:15624-15633.
Školová, B., B. Janůšová, and K. Vávrová. 2016. Ceramides with a pentadecasphingosine
chain and short acyls have strong permeabilization effects on skin and model lipid
membranes. Biochimica et Biophysica Acta - Biomembranes 1858:220-232.
Velkova, V., and M. Lafleur. 2002. Influence of the lipid composition on the organization
of skin lipid model mixtures: An infrared spectroscopy investigation. Chemistry and
Physics of Lipids 117:63-74.
Groen, D., G. S. Gooris, D. J. Barlow, M. J. Lawrence, J. B. van Mechelen, B. Demé, and
J. A. Bouwstra. 2011. Disposition of ceramide in model lipid membranes determined by
neutron diffraction. Biophysical Journal 100:1481-1489.
23. Mojumdar, E. H., D. Groen, G. S. Gooris, D. J. Barlow, M. J. Lawrence, B. Demé, and J.
A. Bouwstra. 2013. Localization of cholesterol and fatty acid in a model lipid membrane:
A neutron diffraction approach. Biophysical Journal 105:911-918.
24. Mojumdar, E. H., Z. Kariman, L. van Kerckhove, G. S. Gooris, and J. A. Bouwstra. 2014.
The role of ceramide chain length distribution on the barrier properties of the skin lipid
membranes. Biochimica et Biophysica Acta (BBA) - Biomembranes 1838:2473-2483.
Akinshina, A., C. Das, and M. G. Noro. 2016. Effect of monoglycerides and fatty acids
on a ceramide bilayer. Physical Chemistry Chemical Physics 18:17446-17460.
Das, C., M. G. Noro, and P. D. Olmsted. 2009. Simulation studies of stratum corneum
lipid mixtures. Biophysical Journal 97:1941-1951.
Das, C., M. G. Noro, and P. D. Olmsted. 2013. Lamellar and Inverse Micellar Structures
of Skin Lipids: Effect of Templating. Physical Review Letters 111:148101-148101.
Das, C., M. G. Noro, and P. D. Olmsted. 2014. Fast cholesterol flip-flop and lack of
swelling in skin lipid multilayers. Soft Matter 10:7346-7352.
Guo, S., T. C. Moore, C. R. Iacovella, L. A. Strickland, and C. McCabe. 2013.
Simulation study of the structure and phase behavior of ceramide bilayers and the role of
lipid head group chemistry. Journal of Chemical Theory and Computation 9:5116-5126.
Gupta, R., B. S. Dwadasi, and B. Rai. 2016. Molecular Dynamics Simulation of Skin
Lipids: Effect of Ceramide Chain Lengths Bilayer Properties. The Journal of Physical
Chemistry B 120:12536-12546.
Gupta, R., and B. Rai. 2015. Molecular Dynamics Simulation Study of Skin Lipids:
Effects of the Molar Ratio of Individual Components over a Wide Temperature Range.
The Journal of Physical Chemistry B 119:11643-11655.
Hoopes, M. I., M. G. Noro, M. L. Longo, and R. Faller. 2011. Bilayer structure and lipid
dynamics in a model stratum corneum with oleic acid. The Journal of Physical Chemistry
B 115:3164-3171.
25.
26.
27.
28.
29.
30.
31.
32.
38
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
Imai, Y., X. Liu, J. Yamagishi, K. Mori, S. Neya, and T. Hoshino. 2010. Computational
analysis of water residence on ceramide and sphingomyelin bilayer membranes. Journal
of Molecular Graphics & Modelling 29:461-469.
Notman, R., J. Anwar, W. J. Briels, M. G. Noro, and W. K. den Otter. 2008. Simulations
of skin barrier function: free energies of hydrophobic and hydrophilic transmembrane
pores in ceramide bilayers. Biophysical Journal 95:4763-4771.
Notman, R., W. K. den Otter, M. G. Noro, W. J. Briels, and J. Anwar. 2007. The
permeability enhancing mechanism of DMSO in ceramide bilayers simulated by
molecular dynamics. Biophysical Journal 93:2056-2068.
Paloncýová, M., R. H. DeVane, B. P. Murch, K. Berka, and M. Otyepka. 2014.
Rationalization of reduced penetration of drugs through ceramide gel phase membrane.
Langmuir 30:13942-13948.
Paloncýová, M., K. Vávrová, Ž. Sovová, R. H. DeVane, M. Otyepka, and K. Berka.
2015. Structural Changes in Ceramide Bilayers Rationalize Increased Permeation through
Stratum Corneum Models with Shorter Acyl Tails. The Journal of Physical Chemistry B
119:9811-9819.
Pandit, S. A., and H. L. Scott. 2006. Molecular-dynamics simulation of a ceramide
bilayer. The Journal of Chemical Physics 124:14708-14708.
Uppulury, K., P. S. Coppock, and J. T. Kindt. 2015. Molecular Simulation of the DPPE
Lipid Bilayer Gel Phase: Coupling Between Molecular Packing Order and Tail Tilt
Angle. The Journal of Physical Chemistry B.
Sparr, E., L. Eriksson, J. A. Bouwstra, and K. Ekelund. 2001. AFM study of lipid
monolayers: III. Phase behavior of ceramides, cholesterol and fatty acids. Langmuir
17:164-172.
Školová, B., K. Hudska, P. Pullmannova, A. Kovacik, K. Palat, J. Roh, J. Fleddermann, I.
Estrela-Lopis, and K. Vávrová. 2014. Different Phase Behavior and Packing of
Ceramides with Long (C16) and Very Long (C24) Acyls in Model Membranes: Infrared
Spectroscopy Using Deuterated Lipids. The Journal of Physical Chemistry B 118:10460-
10470.
Oguri, M., G. S. Gooris, K. Bito, and J. A. Bouwstra. 2014. The effect of the chain length
distribution of free fatty acids on the mixing properties of stratum corneum model
membranes. Biochimica et Biophysica Acta (BBA) - Biomembranes 1838:1851-1861.
Klauda, J. B., R. M. Venable, J. A. Freites, J. W. O'Connor, D. J. Tobias, C. Mondragon-
Ramirez, I. Vorobyov, A. D. MacKerell, and R. W. Pastor. 2010. Update of the
CHARMM all-atom additive force field for lipids: validation on six lipid types. The
Journal of Physical Chemistry B 114:7830-7843.
Gillams, R. J., J. V. Busto, S. Busch, F. M. Goñi, C. D. Lorenz, and S. E. McLain. 2015.
Solvation and hydration of the ceramide headgroup in a non-polar solution. The Journal
of Physical Chemistry B 119:128-139.
Gillams, R. J., C. D. Lorenz, and S. E. McLain. 2016. Comparative atomic-scale
hydration of the ceramide and phosphocholine headgroup in solution and bilayer
environments. The Journal of Chemical Physics 144:225101-225101.
Jorgensen, W. L., J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L. Klein. 1983.
Comparison of simple potential functions for simulating liquid water. The Journal of
Chemical Physics 79:926-926.
39
49.
50.
51.
52.
53.
47. Marinari, E., and G. Parisi. 1992. Simulated Tempering: A New Monte Carlo Scheme.
Europhysics Letters 19:451-458.
48.
Abraham, M. J., T. Murtola, R. Schulz, S. Pall, J. C. Smith, B. Hess, and E. Lindahl.
2015. Gromacs: High performance molecular simulations through multi-level parallelism
from laptops to supercomputers. SoftwareX 1-2:19-25.
Hoover, W. G. 1985. Canonical Dynamics: Equilibrium Phase-Space Distributions.
Physical Review A 31:1695-1967.
Parrinello, M., and A. Rahman. 1981. Polymorphic transitions in single crystals: A new
molecular dynamics method. Journal of Applied Physics 52:7182-7190.
Essmann, U., L. Perera, M. L. Berkowitz, T. Darden, H. Lee, and L. G. Pedersen. 1995.
A smooth particle mesh Ewald method. The Journal of Chemical Physics 103:8577-8593.
Das, C., P. D. Olmsted, and M. G. Noro. 2009. Water permeation through stratum
corneum lipid bilayers from atomistic simulations. Soft Matter 5:4549-4549.
Hartkamp, R., T. C. Moore, C. R. Iacovella, M. A. Thompson, P. Bulsara, D. J. Moore,
and C. McCabe. 2016. Investigating the Structure of Multicomponent Gel Phase Lipid
Bilayers. Biophysical Journal 111:813-823.
Gunstone, F. D., J. L. Harwood, and F. B. Padley. 1994. The Lipid Handbook. Chapman
and Hall.
Boughter, C. T., V. Monje-Galvan, W. Im, and J. B. Klauda. 2016. Influence of
Cholesterol on Phospholipid Bilayer Structure and Dynamics. The Journal of Physical
Chemistry B 120:11761-11772.
Shieh, H. S., L. G. Hoard, and C. E. Nordman. 1977. Crystal structure of anhydrous
cholesterol. Nature 267:287-289.
Tieleman, D. P., and H. J. C. Berendsen. 1996. Molecular dynamics simulations of a fully
hydrated dipalmitoylphosphatidylcholine bilayer with different macroscopic boundary
conditions and parameters. Journal of Chemical Physics 105:4871-4880.
54.
55.
56.
57.
40
|
1704.06685 | 1 | 1704 | 2017-04-21T19:20:00 | A Continuum Poisson-Boltzmann Model for Membrane Channel Proteins | [
"physics.bio-ph"
] | Membrane proteins constitute a large portion of the human proteome and perform a variety of important functions as membrane receptors, transport proteins, enzymes, signaling proteins, and more. The computational studies of membrane proteins are usually much more complicated than those of globular proteins. Here we propose a new continuum model for Poisson-Boltzmann calculations of membrane channel proteins. Major improvements over the existing continuum slab model are as follows: 1) The location and thickness of the slab model are fine-tuned based on explicit-solvent MD simulations. 2) The highly different accessibility in the membrane and water regions are addressed with a two-step, two-probe grid labeling procedure, and 3) The water pores/channels are automatically identified. The new continuum membrane model is optimized (by adjusting the membrane probe, as well as the slab thickness and center) to best reproduce the distributions of buried water molecules in the membrane region as sampled in explicit water simulations. Our optimization also shows that the widely adopted water probe of 1.4 {\AA} for globular proteins is a very reasonable default value for membrane protein simulations. It gives an overall minimum number of inconsistencies between the continuum and explicit representations of water distributions in membrane channel proteins, at least in the water accessible pore/channel regions that we focus on. Finally, we validate the new membrane model by carrying out binding affinity calculations for a potassium channel, and we observe a good agreement with experiment results. | physics.bio-ph | physics | A
Continuum
Poisson--‐Boltzmann
Model
for
Membrane
Channel
Proteins
Li
Xiao1,
Jianxiong
Diao2,
D'Artagnan
Greene2,
Junmei
Wang3,
and
Ray
Luo1,2,4,5
University
of
California,
Irvine,
CA
92697,
Biochemistry,
4.
Chemical
and
Materials
Physics
Graduate
Program,
5.
Department
of
Chemical
Engineering
and
Materials
Science,
1.
Department
of
Biomedical
Engineering,
2.
Department
of
Molecular
Biology
and
3.
Department
of
Pharmaceutical
Sciences,
University
of
Pittsburg,
Pittsburgh,
PA
15261
Membrane
proteins
constitute
a
large
portion
of
the
human
proteome
and
perform
a
variety
of
important
functions
as
membrane
receptors,
transport
proteins,
enzymes,
signaling
proteins,
and
more.
The
computational
studies
of
membrane
proteins
are
usually
much
more
complicated
than
those
of
globular
proteins.
Here
we
propose
a
new
continuum
model
for
Poisson--‐Boltzmann
calculations
of
membrane
channel
proteins.
Major
improvements
over
the
existing
continuum
slab
model
are
as
follows:
1)
The
location
and
thickness
of
the
slab
model
are
fine--‐tuned
based
on
explicit--‐solvent
MD
simulations.
2)
The
highly
different
accessibility
in
the
membrane
and
water
regions
are
addressed
with
a
two--‐
step,
two--‐probe
grid
labeling
procedure,
and
3)
The
water
pores/channels
are
automatically
identified.
The
new
continuum
membrane
model
is
optimized
(by
adjusting
the
membrane
probe,
as
well
as
the
slab
thickness
and
center)
to
best
reproduce
the
distributions
of
buried
water
molecules
in
the
membrane
region
as
sampled
in
explicit
water
simulations.
Our
optimization
also
shows
that
the
widely
adopted
water
probe
of
1.4
Å
for
globular
proteins
is
a
very
reasonable
default
value
for
membrane
protein
simulations.
It
gives
an
overall
minimum
number
of
inconsistencies
between
the
continuum
and
explicit
representations
of
water
distributions
in
membrane
channel
proteins,
at
least
in
the
water
accessible
pore/channel
regions
that
we
focus
on.
Finally,
we
validate
the
new
membrane
model
by
carrying
out
binding
affinity
calculations
for
a
potassium
channel,
and
we
observe
a
good
agreement
with
experiment
results.
1
Please
send
correspondence
to:
[email protected]
Introduction
Membrane
proteins
constitute
a
large
portion
of
the
human
proteome
and
perform
a
variety
of
important
functions,
such
as
membrane
receptors,
transport
proteins,
enzymes,
and
signaling
proteins
1.
These
important
proteins
have
become
primary
drug
targets
in
modern
medicine:
over
60%
of
all
drugs
target
these
proteins
2--‐4.
However,
the
study
of
membrane
proteins
is
usually
much
more
complicated
than
that
of
globular
proteins,
both
experimentally
and
computationally.
For
experimental
studies,
the
difficulty
of
obtaining
a
high--‐resolution
structure
is
an
obstacle,
especially
for
studies
that
involve
proteins
found
in
humans.
For
computational
studies,
modeling
of
the
membrane
environment
is
also
an
important
consideration.
Since
most
biomolecular
systems
exist
in
an
aqueous
environment,
it
is
important
to
account
for
solvent
effects.
There
are
two
ways
to
include
solvent
effects
in
a
computational
simulation:
explicit
and
implicit
solvation.
In
explicit
solvation
modeling,
each
solvent
atom
is
modeled
explicitly.
Although
this
is
the
most
accurate
method,
what
we
are
interested
in
is
often
not
the
properties
of
the
solvent
itself,
but
rather
its
influence
on
the
solute
molecules.
In
addition,
accurately
capturing
the
solvent
influence
in
a
statistically
meaningful
way
requires
sampling
either
from
an
ensemble
of
trajectories
or
from
a
single
very
long
trajectory,
which
is
very
computationally
demanding.
Implicit
solvation
modeling
provides
an
attractive
alternative
wherein
the
solvent
molecules
are
collectively
modeled
as
a
continuum.
In
implicit
solvent
models,
although
the
details
of
individual
solvent
atoms
are
lost,
the
relevant
important
statistically
averaged
effects
can
still
be
preserved
by
design.
Since
solvent
molecules
typically
constitute
the
major
portion
of
molecules
for
an
explicit
solvent
simulation,
implicit
solvent
modeling
can
lead
to
much
2
more
efficient
simulations
5--‐21.
In
addition
to
water,
membrane
molecules
should
also
be
included
when
modeling
solvation
effects,
and
implicit
membrane
modeling
has
also
been
developed
22--‐28.
A
key
issue
in
developing
implicit
solvent
models
is
the
modeling
of
electrostatic
interactions.
The
Poisson--‐Boltzmann
equation
(PBE)
has
been
established
as
a
fundamental
equation
to
model
continuum
electrostatic
interactions
29--‐47.
The
solvent
molecules
are
modeled
as
a
continuum
with
a
high
dielectric
constant,
and
the
solute
atoms
are
modeled
as
a
continuum
with
a
low
dielectric
constant
and
buried
atomic
charges.
The
effect
of
charged
ions
in
the
solvent
region
is
included
by
adding
mobile
charge
density
terms
that
obey
Boltzmann
distributions.
The
potential
of
the
full
system
is
then
governed
by
the
partial
differential
equation:
where
∇
is
the
spatial
gradient
operator,
ε
is
the
dielectric
constant
distribution,
𝜙
is
the
electrostatic
potential
distribution,
𝜌!
is
the
charge
density
of
the
solute
(usually
modeled
as
a
set
of
discrete
point
charges),
ci
is
the
concentration
of
the
ith
solvent
ion
species
in
bulk,
e
is
the
absolute
charge
of
an
electron,
zi
is
the
valence
for
the
ith
ion,
kB
is
Boltzmann's
constant,
T
is
the
temperature,
and
λ
is
the
Stern
layer
masking
function,
which
is
0
within
or
1
outside
of
the
Stern
layer.
The
PBE
is
a
non--‐linear
elliptical
partial
differential
equation.
There
is
no
closed
form
solution,
and
thus,
numerical
methods
are
often
required
for
biomolecular
applications
22,
36,
44,
45,
48--‐87.
Efficient
numerical
PBE--‐based
solvent
models
have
been
widely
used
to
study
biological
processes
including
predicting
pKa
values
88--‐91,
computing
solvation
and
binding
free
energies
92--‐101,
and
protein
folding
102--‐112.
Predicting
protein--‐
∇⋅ε∇φ= −4πρ0 − 4π eziciλexp(−eziφ/ kBT)
3
(1)
i∑
ligand
binding
affinities
is
one
of
the
major
applications
for
implicit
solvent
free
energy
calculations.
In
the
Amber
software
package,
MMPBSA
is
the
module
performing
such
calculations
113--‐118.
Implicit
membrane
modeling
has
also
been
applied
and
developed
in
binding
free
energy
calculations.
There
are
a
noticeable
number
of
pioneer
works
that
implement
implicit
membrane
modeling
in
several
PB
packages,
such
as
APBS
27,
Delphi
28,
119,
PBEQ
78,
120,
and
PBSA
121--‐123.
All
of
them
add
the
membrane
as
a
slab
with
a
relatively
low
dielectric
constant
that
is
embedded
in
water
for
PBE
calculations.
Our
previous
work
implemented
an
implicit
membrane
model
into
the
PBE
framework
121.
The
implicit
membrane
model
can
be
readily
interfaced
with
the
existing
MMPBSA
program
113--‐118
to
perform
binding
free
energy
calculations
of
several
protein
structures
embedded
in
a
membrane
87,
124.
However,
a
problem
arises
when
those
membrane
proteins
contain
a
pore
or
a
gated
channel,
since
the
region
of
the
channel
is
usually
permeable
and
should
be
composed
of
water.
Therefore,
a
simple
slab--‐like
membrane
setup
may
cause
problems
if
the
membrane
protein
contains
pore--‐
or
channel--‐
like
region(s).
Similar
to
the
approaches
adopted
in
the
community
28,
119,
125,
we
dealt
with
this
issue
by
manually
defining
the
pore
region
as
a
cylinder,
and
we
then
set
the
dielectric
constant
within
the
cylindrical
region
as
that
of
water
if
it
was
not
occupied
by
protein
atoms.
The
limitation
of
this
method
is
that,
for
every
snapshot
of
a
trajectory,
we
need
to
visualize
and
locate
the
cylinder
by
hand,
which
is
neither
efficient
nor
practical
given
the
large
number
of
snapshots
that
must
be
processed
for
converged
calculations.
In
this
work,
we
propose
a
new
continuum
membrane
model
for
PBE
calculations
of
biomolecules.
Major
improvements
from
the
existing
continuum
slab
model
are
the
following:
1)
an
explicit
solvent
MD
simulation
was
exploited
to
fine
tune
the
slab
model,
4
i.e.
its
exact
location
and
thickness,
to
best
reproduce
the
solvent
accessibility
and
the
water
accessible
channel,
2)
a
two--‐step,
two--‐probe
initial
grid
labeling
procedure
was
adopted
to
address
highly
different
accessibility
in
the
membrane
region
and
water
region,
and
3)
a
depth--‐first
search
algorithm
was
introduced
to
detect
the
water
pores/channels
automatically
based
on
the
initial
grid
labels.
This
procedure
follows
our
basic
algorithm
proposed
for
globular
proteins,
and
adds
little
overall
overhead
in
the
application
of
linear
finite--‐difference
PBE
solvers
to
typical
membrane
proteins.
Methods
The
Poisson--‐Boltzmann
equation
(Eqn
(1))
is
widely
used
in
capturing
electrostatic
energy
and
forces
in
implicit
solvent
modeling.
For
systems
with
dilute
ion
concentrations,
the
second
term
on
the
right--‐hand
side
is
usually
linearized,
giving
the
simpler
form:
(2)
where
κ2 = 4π ciei
The
finite--‐difference
method
22,
43,
71--‐83,
86
is
one
of
the
most
popular
methods
used
in
the
numerical
implementation
of
the
PBE.
In
a
typical
procedure,
the
rectangular
grid
covering
the
solution
system
is
first
defined.
Next,
the
atomic
point
charges
are
mapped
onto
the
grid
points
with
a
predefined
assignment
function.
Third,
the
dielectric
constant
distribution
is
mapped
to
the
grid
edges.
The
discretized
linear
system
is
then
turned
to
a
linear
solver
to
solve
for
potentials
on
the
grid
points,
which
can
be
expressed
as:
∇ ⋅ε∇φ= −4πρ0 +λκ2φ
.
2 / kBT
2zi
i∑
5
⎡
⎢
⎢
⎢
⎢
⎢
⎢
⎢
⎣
4πρ(i, j,k)
h
εx,εy,
εz
−
⎤
⎥
⎥
⎥
⎥
⎥
⎥
⎥
⎦
+ h2λ(i, j,k)κ2φ(i, j,k) = −
εx(i, j,k)φ(i +1, j,k) +εx(i −1, j,k)φ(i −1, j,k)
⎤
⎡
⎥
⎢
εy(i, j,k)φ(i, j +1,k) +εy(i, j −1,k)φ(i, j −1,k)
⎥
⎢
εz(i, j,k)φ(i, j,k +1) +εz(i, j,k −1)φ(i, j,k −1)
⎥
⎢
⎦
⎣
εx(i −1, j,k) +εx(i, j,k) +εy(i, j −1,k)
⎤
⎡
⎥φ(i, j,k)
⎢
εy(i, j,k) +εz(i, j,k −1) +εz(i, j,k)
⎦
⎣
(3)
and
Here
represent
the
dielectric
constants
for
grid
edges
along
the
x,
y,
and
z
directions
respectively,
and
h
represents
the
grid
spacing.
This
study
focuses
on
how
to
set
up
linear
PBE
applications
for
membrane
systems.
A
major
issue
is
the
presence
of
the
membrane
and
its
influence
on
the
dielectric
constant
distribution.
In
globular
proteins,
the
solvent
excluded
surface
(SES)43,
126--‐130
is
often
used
as
a
boundary
separating
the
high
dielectric
water
exterior
and
the
low
dielectric
protein
interior.
The
presence
of
the
membrane
introduces
at
least
a
third
region.
In
this
study,
we
adopt
the
uniform
membrane
dielectric
model,
though
our
procedure
can
be
easily
extended
to
accommodate
another
often
used
depth--‐dependent
membrane
dielectric
model.
The
first
step
is
to
introduce
a
membrane
region
to
the
existing
solvent
excluded
surface
procedure
with
minimum
invasion
to
the
program
and
minimum
efficiency
lost.
The
SES
is
the
most
common
surface
definition
used
to
describe
the
dielectric
interface
between
the
two
piece--‐wise
dielectric
constants.
In
fact,
comparative
analysis
of
PB--‐based
solvent
models
and
TIP3P
solvent
models
have
shown
that
the
SES
definition
is
reasonable
in
the
calculation
of
reaction
field
energies
and
electrostatic
potentials
of
mean
force
fields.
131--‐133
Here,
we
follow
the
idea
from
Rocchia
et
al.
130
and
Wang
et
al.
43
of
mapping
the
SES
to
a
finite--‐difference
grid.
While
keeping
the
variables
used
to
label
the
solvent
and
solute
regions,
we
also
introduce
a
new
variable
to
label
the
membrane
region.
Considering
the
6
membrane
molecules
are
usually
larger
than
solvent
molecules,
we
use
two
different
solvent
probe
radii
to
set
up
the
membrane
and
solvent
regions.
And
finally,
we
assign
the
dielectric
constant
on
each
region
and
the
interface.
Grid
point
labeling
Our
general
strategy
is
to
model
the
membrane
as
a
second
continuum
solvent
of
finite
region,
i.e.
a
slab
located
at
a
user
specified
position.
The
essence
of
the
algorithm
is
to
determine
both
the
membrane
accessibility
and
water
accessibility
around
a
molecular
solute.
Assisted
with
both
sets
of
accessibility
data,
the
presence
of
water
channels
or
water
pores
within
the
membrane
region
can
then
be
identified
in
the
next
step.
Due
to
the
much
larger
size
of
lipid
molecules,
a
separate
solvent
probe
(mprob)
must
be
used
to
determine
the
membrane
accessibility.
This
is
apparently
much
larger
than
the
water
probe
(dprob).
The
influence
of
both
probes
on
reproducing
the
solvent
accessible
surface
of
a
membrane
protein
is
presented
in
Results
and
Discussion.
In
Amber/PBSA,
an
integer
array
insas
is
used
to
label
whether
the
grid
point
is
outside
the
solute
region
(insas<0)
or
inside
the
solute
region
(insas>0)
for
fast
mapping
of
solvent
accessibility
information
43.
This
labeling
scheme
has
been
extended
to
map
all
commonly
used
surfaces,
SES,
SAS,
VDW,
and
DEN
in
recent
Amber
and
AmberTools
releases.36,
43,
80,
82,
87,
134,
135
To
minimize
the
interference
to
existing
procedures
and
maximize
efficiency,
a
separate
integer
array
inmem
is
used
to
label
whether
a
grid
point
is
inside
the
membrane
(inmem>0)
or
outside
the
membrane
(inmem=0).
Specifically,
the
grid
labeling
algorithm
can
be
summarized
as
the
following
five
steps:
7
0. Initialize insas of all grid points as "-4", i.e. in the bulk
solvent and salt region, and inmem of all grid points as "0",
i.e. outside the membrane region.
1. Using mprob as the solvent probe radius, label insas of all
grid points as "-3" if within the stern layer; "-2" if within
the solvent accessible surface layer; "-1" if within the
reentry region but outside the SES; "1" if within the reentry
region but inside the SES; "2" if inside the VDW surface.
2. Add a slab perpendicular to the z-axis as the membrane region
at the specified location. Label inmem of the membrane-region
grid points with insas<0 as "1".
3. Apply the depth-first search algorithm to detect any possible
membrane accessible grid point that is not connected to the
bulk membrane. If so, relabel its inmem as "0".
4. For each grid point with (inmem=0) within the slab, if it has
a neighbor with (inmem=1) within the distance cutoff of
memmaxd, relabel its inmem as "2".
5. Using dprob as the solvent probe radius, relabel insas of all
grid points as "-3" if within the stern layer; "-2" if within
the solvent accessible surface layer; "-1" if within the
reentry region outside the SES; "1" if within the reentry
region inside the SES; "2" if inside the VDW surface.
A
few
explanations
are
in
order
here.
First,
inmem
is
determined
in
Step
2
through
Step
4,
so
that
its
value
is
controlled
by
both
the
mprob--‐generated
insas
and
the
depth--‐first
search
algorithm.
Second,
a
new
variable
(memmaxd)
is
introduced
in
Step
4.
Since
mprob
is
usually
much
larger
than
dprob,
there
exists
a
thin
layer
of
grid
points
with
insas>0
and
inmem=0
between
the
membrane
region
and
the
protein
region.
If
the
grid
labels
are
set
this
way,
these
grid
points
would
be
labeled
as
water
in
a
later
processing
stage
of
our
method,
thus
leading
to
an
artificial
layer
of
water
between
the
protein
and
membrane.
To
resolve
this
issue,
a
cutoff
distance
of
memmaxd
is
introduced
to
represent
the
maximum
difference
between
the
SES
surfaces
generated
by
mprob
and
dprob.
This
is
estimated
to
be
mprob−dprob
assuming
maximum
reentry
by
dprob.
Thus
Step
4
changes
the
inmem
labels
of
the
grid
points
from
0
(mprob
inaccessible)
to
2
(mprob
accessible)
if
they
are
memmaxd
inside
the
mprob--‐generated
SES.
The
correction
effectively
removes
the
artificial
layer
of
water
between
the
protein
and
the
membrane.
Here
the
revised
inmem
values
are
8
set
to
be
"2"
so
these
grid
points
would
not
interfere
with
the
subsequent
search.
Note
too
that
this
correction
does
not
change
the
protein
interior
definition,
which
is
defined
with
the
water
dprob.
Nevertheless,
it
does
have
the
effect
of
pushing
back
the
potential
buried
water
pockets,
if
any,
from
the
protein--‐membrane
interface.
In
summary,
the
three
different
regions
that
are
readily
available
for
further
processing
after
the
grid--‐labeling
step
are:
1. Solute region: insas(i,j,k)>0
2. Membrane region: insas(i,j,k)<0 and inmem(i,j,k)>0
3. Solvent region: insas(i,j,k)<0 and inmem(i,j,k)=0
Membrane
pore/channel
detection
Step
3
in
the
above
general
grid--‐labeling
algorithm
is
meant
to
identify
pore--‐
or
channel--‐
like
water--‐accessible
water
pockets
within
a
user--‐specified
membrane
region.
Given
the
convention
that
the
membrane
is
parallel
with
the
xy
plane,
the
membrane
region
can
be
mathematically
defined
to
be
all
grid
points
within
[zmin, zmax]. Thus the method starts
by initializing all grid points that are defined as solvent (insas<0) within [zmin, zmax] as
inmem=1. Next the recursive depth-first search algorithm is used to traverse all grid points to
see whether they are connected or not. Our goal of using the algorithm is to walk and label
recursively all grid points in the non-protein regions within [zmin, zmax]. Upon completion,
all grid points that are not connected to the membrane region (i.e. the pore region) are labeled
back as the water region (inmem=0). To facilitate the bookkeeping of the search, the variable
kzone is introduced to label the different regions: the protein region (kzone=0), the membrane
region (kzone=1), and the water regions (kzone>1). Since the search starts from the edge of
the membrane slab, the first region found is always the membrane region (kzone=1), and the
rest are the water regions or the protein region. In general multiple kzone values are assigned
9
because most water-accessible regions are not connected. The algorithm can be summarized as
shown below:
nzone = 0; kzone = -1
for k = zmin:zmax
for j,i = 1:n
cycle
if kzone(i,j,k) != -1 then
end if
if insas(i,j,k) > 0 then
else
kzone(i,j,k) = 0
nzone = nzone + 1
kzone(i,j,k) = nzone
call walk(i,j,k,kzone,nzone)
end if
end
end
recursive subroutine walk(i,j,k,kzone,nzone)
kzone(i,j,k) = nzone
if (kzone(i+1,j,k) == -1 .and. insas(i+1,j,k)<0)
call walk(i+1,j,k,kzone,nzone)
if (kzone(i-1,j,k) == -1 .and. insas(i-1,j,k)<0)
call walk(i-1,j,k,kzone,nzone)
if (kzone(i,j+1,k) == -1 .and. insas(i,j+1,k)<0)
call walk(i,j+1,k,kzone,nzone)
if (kzone(i,j-1,k) == -1 .and. insas(i,j-1,k)<0)
call walk(i,j-1,k,kzone,nzone)
if (k+1<=zmax .and. kzone(i,j,k+1) == -1 .and. insas(i,j,k+1)<0)
call walk(i,j,k+1,kzone,nzone)
if (k-1>=zmin .and. kzone(i,j,k-1) == -1 .and. insas(i,j,k-1)<0)
call walk(i,j,k-1,kzone,nzone)
In
this
way,
all
grid
points
with
kzone>1
are
water
accessible,
and
inmem
of
these
end recursive subroutine walk
grid
points
are
set
back
to
0,
i.e.
membrane
inaccessible.
Mapping
solvent/membrane
accessibility
to
dielectric
constants
10
In
this
study,
we
adopted
a
three--‐dielectric
model
to
model
the
membrane--‐protein
electrostatics.
The
dielectric
constants
for
the
three
different
regions
are
denoted
as
inε
outε (solvent)
and
εmem
(membrane),
respectively.
(solute),
The
next
step
is
to
map
the
grid
labeling
information
into
the
dielectric
constants
at
the
midpoints
on
all
grid
edges.
The
general
principle,
to
be
consistent
with
Wang
et
al,
43
is
that
the
dielectric
constant
of
a
grid
edge
should
be
equal
to
the
dielectric
constant
in
the
region
where
the
two
flanking
grid
points
reside.
When
the
two
neighboring
grid
points
belong
to
different
dielectric
regions,
the
weighted
harmonic
averaging
(WHA)
method
is
used
to
calculate
the
"fractional"
dielectric
constant
based
on
the
precise
intersection
point
where
the
molecular
surface
cut
the
grid
edge
43.
Specifically
the
dielectric
constant
is
assigned
as:
(4)
where
a
denotes
the
fraction
of
the
grid
edge
in
region
1.
Eqn
(4)
is
applied
on
three
different
kinds
of
interfaces:
(5)
The
assignment
of
dielectric
constants
on
the
solute
and
solvent
interface
is
the
same
as
Wang
et
al
43.
We
now
consider
the
grid
edge
between
(i,j,k)
and
(i+1,j,k);
the
grid
edge
can
be
classified
according
to
the
rules
in
Table
I.
The
procedure
of
assigning
the
dielectric
constants
on
the
membrane
related
region
and
interface
is
as
follows,
for
each
of
the
x--‐,
y--‐,
and
z--‐edges,
respectively.
ε1 = εin,ε2 = εout solute and solvent interface
ε1 = εin,ε2 = εmem solute and membrane interface
ε1 = εout,ε2 = εmem solvent and membrane interface
11
,
1
1− a
ε2
+
a
ε1
ε=
insas(i,j,k)
>0
>0
>0
>0
>0
>0
>0
>0
<0
<0
<0
<0
<0
<0
<0
<0
insas(i+1,j,k)
>0
>0
>0
>0
<0
<0
<0
<0
>0
>0
>0
>0
<0
<0
<0
<0
inmem(i,j,k)
>0
>0
=0
=0
>0
>0
=0
=0
>0
>0
=0
=0
>0
>0
=0
=0
inmem(i+1,j,k)
>0
=0
>0
=0
>0
=0
>0
=0
>0
=0
>0
=0
>0
=0
>0
=0
region
inside solute
inside solute
inside solute
inside solute
solute and membrane interface
solute and solvent interface
solute and membrane interface
solute and solvent interface
solute and membrane interface
solute and membrane interface
solute and solvent interface
solute and solvent interface
inside membrane
inside solvent
inside solvent
inside solvent
Table I: Different edges of dielectric constants defined by adjacent values of insas and inmem.
For
x--‐edges,
fractional
membrane
edges
are
only
possible
with
the
membrane--‐
solute
interface,
so
that
the
following
pseudo
code
can
be
added
to
the
existing
dielectric
mapping
procedure:
If (inmem(i,j,k)>0 .or. inmem(i+1,j,k)>0) then
If (insas(i,j,k)>0 .and. insas(i+1,j,k)>0) then
εx(i, j,k) =εin
// grid edge in solute
else if ((insas(i,j,k)>0 .and. inmem(i+1,j,k)>0) .or.
(insas(i+1,j,k)>0 .and. inmem(i,j,k)>0) ) then
// grid edge between membrane and solute
else if (insas(i,j,k)>0 .or. insas(i+1,j,k)>0) then
εx(i, j,k) =
εx(i, j,k) =
1
1− a
+
εmem
a
εin
1
1− a
+
εout
a
εin
// grid edge between solvent and solute
else if(inmem(i,j,k)>0 .and. inmem(i+1,j,k)>0) then
εx(i, j,k) = εmem
// grid edge in membrane
end if
12
end if
Here
a
is
the
fraction
of
grid
edge
in
the
solute
region.
The
algorithm
along
the
y--‐axis
is
similar
to
the
x--‐axis,
as
follows:
If (inmem(i,j,k)>0 .or. inmem(i,j+1,k)>0) then
If (insas(i,j,k)>0 .and. insas(i,j+1,k)>0) then
εy(i, j,k) = εin
// grid edge in solute
else if ((insas(i,j,k)>0 .and. inmem(i,j+1,k)>0) .or.
(insas(i,j+1,k)>0 .and. inmem(i,j,k)>0) ) then
// grid edge between membrane and solute
else if (insas(i,j,k)>0 .or. insas(i,j+1,k)>0) then
εy(i, j,k) =
εy(i, j,k) =
1
1− a
+
εmem
a
εin
1
1− a
+
εout
a
εin
// grid edge between solvent and solute
else if(inmem(i,j,k)>0 .and. inmem(i,j+1,k)>0) then
εy(i, j,k) = εmem
end if
// grid edge in membrane
end if
For
the
dielectric
constant
mapping
along
the
z--‐axis,
the
algorithm
also
involves
the
solvent--‐membrane
interface;
the
algorithm
should
also
take
care
of
this,
as
follows:
If (inmem(i,j,k)>0 .or. inmem(i,j,k+1)>0)
If (insas(i,j,k)>0 .and. insas(i,j,k+1)>0) then
εz(i, j,k) =εin
else if ((insas(i,j,k)>0 .and. inmem(i,j,k+1)>0) .or.
(insas(i,j,k+1)>0 .and. inmem(i,j,k)>0) ) then
εz(i, j,k) =
// grid edge between membrane and solute
else if (insas(i,j,k)>0 .or. insas(i,j,k+1)>0) then
1
1− a
+
εmem
a
εin
1
1− a
+
εout
a
εin
εz(i, j,k) =
// grid edge between solvent and solute
else if(inmem(i,j,k)>0 .and. inmem(i,j,k+1)>0) then
13
εz(i, j,k) = εmem
1
+
εz(i, j,k) =
a
εout
1− a
εmem
else if (grid edge is cross the slab)
end if
// grid edge cross the slab
// a is fraction of edge in solvent
end if
Finally,
all
the
edges
in
the
water
are
assigned
the
dielectric
constant
of
water,
all
the
edges
in
the
membrane
are
assigned
the
dielectric
constant
of
membrane,
and
all
the
edges
in
the
protein
interior
are
assigned
the
dielectric
constant
of
protein.
For
all
the
edges
crossing
different
regions,
i.e.
between
any
two
of
water,
membrane,
or
protein,
weighted
harmonic
averages
between
the
two
corresponding
dielectric
constants
are
assigned.
Protein
and
complex
structure
preparation
To
calibrate
the
new
continuum
membrane
model
for
channel
detection,
we
simulated
three
channel
proteins
with
crystal
structures:
1K4C
136,
a
KcsA
potassium
channel;
5CFB
137,
an
alpha1
GlyR
Glycine
receptor;
and
5HCJ
138,
a
prokaryotic
pentameric
ligand--‐gated
ion
channel.
To
demonstrate
the
feasibility
of
the
new
continuum
membrane
model
in
data
intensive
binding
affinity
calculations,
we
chose
the
hERG
K+
channel
protein,
given
its
importance
in
drug
discovery
and
the
availability
of
high--‐quality
experimental
data
139.
A
homology
model
of
hERG
K+
channel
was
built
based
on
the
X--‐ray
crystal
structure
of
KcsA
136
(PDB
ID:
1K4C)
using
MODELLER
140
(version
9.15)
with
the
default
settings.
The
amino
acid
sequence
of
hERG
K+
channel
was
directly
extracted
from
the
Swiss--‐Prot
14
database
141
(accession
number:
Q12809
and
entry:
KCNH2_HUMAN).
Sequence
alignment
was
generated
using
CLUSTALX
142
(version
2.1),
showing
a
good
match
in
helices
S5,
S6,
and
the
pore
region,
with
identity
about
44%
(Figure
1).
After
automatic
model
building
and
loop
refinement,
candidate
models
were
evaluated
based
on
the
DOPE
score
from
MODELLER
140.
The
final
homology
model
of
the
hERG
K+
channel
is
shown
in
Figure
2,
which
is
found
to
be
highly
consistent
with
a
previously
reported
model
based
on
a
different
procedure
143--‐146.
Figure 1: Sequence alignment of KcsA and hERG by ClustalX version 2.1. The identified S5
helix, S6 helix, amphipathic helix and pore helix are labeled above the sequence. Asterisks (*):
conserved amino acid residues; colons (:): conserved substitutions; dots (.): semi-conserved
substitutions.
15
Figure 2: Comparison of target and parent structures, showing the secondary structure elements
in homology models of hEGH (red) and KcsA (blue). The plot shows three orientations of the
aligned structure. Top: side view with the binding pocket on the top. Bottom left: viewed from
the binding pocket/extracellular side. Bottom right: viewed from the intracellular side.
Initial
complex
structures
of
the
hERG
K+
channel
with
its
inhibitors
were
generated
with
the
SURFLEX--‐DOCK
program
in
Sybyl--‐X
(version
1.3).
Ten
different
inhibitors
with
experimental
binding
affinities
139
were
chosen
to
assess
the
quality
of
the
MMPBSA
procedure,
including
astemizole
(AST),
sertindole
(SER),
pimozide
(PIM),
droperidol
16
(DRO),
terfenadine
(TE0,
TE1),
domperidone
(DOM),
loratadine
(LOR),
mizolaatine
(MIZ),
perhexiline
(PE0,
PE1)
and
amitriptyline
(AMI).
The
terfenadine
and
perhexiline
are
chiral
molecules
with
two
enantiomers,
so
both
enantiomers
were
used
in
the
docking.
Molecular
dynamics
simulation
MMPBSA
calculations
of
binding
affinities
The
protein
was
first
inserted
into
a
membrane
layer
using
the
CHARMM--‐GUI
lipid
builder
147--‐151.
Lipid
DPPC
was
used
for
the
membrane
layer
with
a
lipid
to
water
ratio
of
29.
The
solvated
membrane
system
first
underwent
a
10,000--‐step
energy
minimization
using
a
5,000--‐step
steepest
descent
followed
by
a
5,000--‐step
conjugate
gradient.
The
main
chain
atoms
for
the
protein
were
then
restrained
with
a
force
constant
of
2
kcal/mol--‐Å2.
Subsequently,
a
5
ps
MD
simulation
was
conducted
to
heat
the
system
from
0
to
100K
followed
by
a
100
ps
MD
simulation
to
heat
the
system
from
100K
to
310K.
This
was
then
followed
with
a
5
ns
simulation
for
equilibration.
Finally,
production
MD
was
run
for
50
ns.
Binding
free
energies
were
computed
using
a
revised
MMPBSA
module124
of
Amber
16
or
AmberTools
2016
134,
135,
152.
The
production
run
trajectory
was
post--‐processed
with
CPPTRAJ
153
in
order
to
remove
the
solvent,
membrane,
and
counter
ions
from
the
receptor--‐ligand
complex.
Snapshots
from
the
last
10
ns
of
the
production
run
were
processed
to
compute
molecular
mechanics
potential
energies
and
solvation
free
energies
in
the
MMPBSA
procedure.
The
binding
free
energy
for
the
protein--‐ligand
complex
was
computed
as
the
difference
between
the
complex
free
energy
and
the
sum
of
the
receptor
and
ligand
free
energies,
as
outlined
in
our
previous
work
124.
The
electrostatic
solvation
17
free
energies
were
calculated
using
the
linearized
PBE
model
as
implemented
in
PBSA
36,
43,
80,
82,
87.
The
non--‐electrostatic
solvation
free
energies
were
calculated
using
either
the
classical
model
or
the
modern
model
as
documented
previously
154.
Additional
computational
details
In
each
PBSA
calculation,
a
finite--‐difference
grid
spacing
of
0.5
Å
was
used
for
MMPBSA
calculations,
which
was
found
to
be
sufficient
due
to
MD
sampling
and
the
approximate
nature
of
the
binding
affinity
calculation
118.
Production
snapshots
up
to
10ns
were
found
to
be
sufficient
to
converge
the
averaging
process
used
in
MMPBSA
calculations
of
these
membrane
protein--‐ligand
complexes.
The
periodic
geometric
multigrid
solver
option
was
employed
with
a
convergence
threshold
of
1.0
x
10--‐3,
and
electrostatic
focusing
was
turned
off
due
to
the
presence
of
the
membrane
87.
The
use
of
a
periodic
boundary
also
allowed
a
somewhat
small
fillratio
(i.e.
the
ratio
of
the
finite--‐difference
box
dimension
over
the
solute
dimension)
of
1.25
to
be
used
in
these
calculations
37.
The
solvation
system
physical
constants
were
set
up
as
follows.
The
membrane
was
modeled
as
a
solid
slab
as
simulated
in
the
explicit
water
MD
trajectories.
The
water
relative
dielectric
constant
was
set
at
80.0.
The
membrane
dielectric
constant
was
set
to
be
7.0
124.
And
the
protein
dielectric
constant
was
set
to
be
20.0
due
to
the
presence
of
charged
ligand
molecules
118,
124.
The
water
phase
ionic
strength
was
set
to
be
150
mM.
The
lower
dielectric
region
within
the
molecular
solutes
was
defined
with
the
classical
solvent
excluded
surface
model
using
a
water
solvent
probe
and
a
membrane
solvent
probe
to
be
optimized
as
described
in
Results
and
Discussion.
The
default
weighted
harmonic
averaging
was
employed
to
assign
18
dielectric
constants
for
boundary
grid
edges
to
reduce
grid
dependency
43.
Charges
and
radii
were
assigned
as
in
the
simulation
topology
files.
Results
and
Discussion
Optimization
of
the
new
slab
membrane
model
Given
the
automatic
procedure
in
place
to
identify
water
channels/pores
with
the
depth--‐first
search
method,
we
further
optimized
the
membrane
probe
value
and
the
slab
membrane
model
(i.e.
its
thickness)
to
best
reproduce
the
distributions
of
buried
water
molecules
in
the
membrane
region
as
sampled
in
explicit
water
MD
simulations.
Three
different
membrane
proteins
with
channels
were
utilized
in
this
optimization:
1K4C,
5HCJ,
and
5CFB.
Three
different
slab
definitions
were
evaluated
to
set
up
the
continuum
membrane
model,
i.e.
the
inner
and
outer
faces
are
chosen
to
be
positioned
at
(1)
the
average
z--‐
coordinates
of
nitrogen
atoms
of
the
lipid
head
groups;
(2)
the
average
z--‐
coordinates
of
the
phosphorus
atoms
of
the
lipid
head
groups;
(3)
the
average
z--‐
coordinates
of
both
nitrogen
and
phosphorus
atoms
in
the
lipid
head
groups.
Here
the
average
z--‐coordinates
are
computed
from
the
explicit--‐water
MD
simulations.
Next,
mprob
values
were
scanned
from
1.4
Å
upwards
to
3.0
Å
with
an
increment
of
0.1
Å.
The
smallest
mprob
value
with
which
these
known
channels
can
be
displayed
was
recorded
as
the
mprob
threshold
in
Table
II
for
all
three
slab
membrane
definitions.
It
should
be
pointed
out
that
a
small
mprob
produces
excessive
membrane
accessibility
in
the
protein
interior
so
that
it
is
more
likely
for
the
buried
membrane
pockets
to
be
19
connected
to
the
bulk
membrane.
Excessive
membrane
accessibility
can
also
be
lessened
by
reducing
the
membrane
thickness,
as
in
the
use
of
phosphorus
atoms
to
define
the
boundaries
of
the
continuum
membrane.
Indeed,
our
analysis
showed
this
setup
caused
the
least
penetration
of
the
continuum
membrane
into
the
protein
interior,
so
the
smallest
mprob
(2.7
Å)
was
needed
to
capture
the
water
channels/pores
for
all
three
tested
proteins.
Figure
3
shows
the
rendering
of
water--‐channels/pores
of
the
three
tested
membrane
proteins
with
the
optimized
mprob.
The
advantage
of
the
optimal
mprob
over
the
default
solvent
probe
of
1.4
Å
is
apparent
by
comparing
the
renderings
generated
with
the
two
probes.
For
all
the
channel
proteins,
the
new
model
automatically
detects
the
water
channels/pores.
Figure
4
further
shows
the
benefit
of
the
depth--‐first--‐search
feature
that
is
a
must
in
the
new
slab
membrane
model.
Without
it,
it
is
apparent
that
none
of
the
water
channels/pores
can
be
identified
for
any
of
the
tested
proteins,
even
using
the
larger
probe
for
the
membrane
region.
Protein
mthick (Å) mcenter(Å) mprob (Å)
1K4C
5CFB
5HCJ
1K4C
5CFB
5HCJ
1K4C
5CFB
5HCJ
mthick=N+–N–
-1.10
64.95
68.60
mthick=P+–P–
-0.97
64.97
68.40
mthick=N+P+– N–P–
-1.04
64.96
68.50
>2.2
>1.7
>3.0
>2.2
>1.6
>2.7
>2.2
>1.7
>3.0
39.24
40.72
41.11
36.13
37.27
37.89
37.69
39.00
39.50
20
Table II The thickness of membrane and mprob thresholds based on the different criterion
measured from MD simulations. Here the mprob threshold is the minimum value with which the
channel is visible with the SES approach. (Top) mthick=N+–N–: The thickness of the membrane
slab is defined as the z-distance between the average head group nitrogen atoms of the lipid
molecules. (Middle) mthick=P+–P–: The thickness of the membrane slab is defined as the z-
distance between the average head group phosphorus atoms of the lipid molecules. (Bottom)
mthick=N+P+– N–P–: the membrane slab is defined as the z-distance between the average head
group centers (i.e. the means of nitrogen and phosphorus atoms) of the lipid molecules. The
membrane center locations were then computed as the mean of the upper and lower bounds.
21
Figure 3 Solvent-solute interface determined with the new continuum membrane model. Left:
mprob is set to be 1.4 Angstroms, the default value of the solvent probe. Right: mprob is set to be
2.7 Angstroms, the optimized value of the membrane probe. Three proteins are tested: 1K4C
(top); 5CFB (middle); 5HCJ (bottom).
22
Figure 4 Same as Figure 3, except without turning on the depth-first search in the pore region
detection. Three proteins are tested: 1K4C (top); 5CFB (middle); 5HCJ (bottom).
Impact
of
the
water
solvent
probe
upon
agreement
with
an
explicit
solvent
simulation
It
is
worth
pointing
out
that
the
agreement
of
the
continuum
membrane
model
also
depends
on
how
we
model
the
water
accessible
region.
The
standard
practice
has
been
to
consider
the
finite
size
of
the
water
molecule
with
a
predefined
probe
radius,
often
taken
as
1.4
Å.
The
probe
is
then
used
to
compute
the
solvent
excluded
surface
used
as
the
interface
separating
the
protein
interior
from
the
water
region.
It
is
apparent
that
the
size
of
the
water
accessible
pores/channels
would
depend
on
how
large
the
water
probe
is
defined.
Thus,
it
is
interesting
to
analyze
how
well
the
widely
used
water
probe
performs
in
the
context
of
membrane
channel
proteins.
This
analysis
was
conducted
in
the
following
manner.
The
distributions
of
water
molecules
(in
the
water
pore/channel
regions)
in
explicit
water
MD
simulations
were
sampled
every
50
ps
over
the
course
of
a
5
ns
production
run.
Note
that
the
protein
atoms
were
all
restrained
to
the
reference
structure
after
equilibration
since
the
focus
was
on
the
water
distribution.
A
total
of
100
frames
worth
of
water
sampling
were
collected
and
were
combined
into
one
snapshot
for
visualization.
This
water
distribution
map
was
used
as
a
reference
to
evaluate
how
the
hard
sphere
SES
surface
behaves
with
one
single
adjustable
parameter,
i.e.
the
water
solvent
probe
(dprob
in
Amber/PBSA).
The
same
three
membrane
channel
proteins
were
analyzed
to
address
this
question.
Specifically,
the
counts
for
the
following
disagreements/mismatches
were
recorded:
(1)
the
absence
of
explicit
water
molecules
in
the
continuum
water
accessible
regions;
and
(2)
the
23
presence
of
explicit
water
molecules
in
the
continuum
water
inaccessible
regions.
The
overall
summary
of
both
mismatches
is
reported
in
Table
III.
Sample
mismatches
are
shown
in
Figure
5.
It
is
interesting
to
note
that
the
standard
value
of
the
water
solvent
probe
of
1.4
Å
is
a
very
reasonable
default
value,
which
gives
an
overall
minimum
number
inconsistencies
between
the
continuum
and
explicit
representations
of
water
of
distributions
in
the
tested
membrane
channel
proteins,
at
least
in
the
water
accessible
pore/channel
regions
that
we
have
focused
on.
Protein
dprob (Å)
No. solvent region
w/o water molecules
No. water molecules in
non-solvent region
Table III Discrepancies in the solvent accessible region between explicit water MD simulations
and the membrane PBSA calculations. Two types of discrepancies were recorded: (1) how many
continuum solvent pockets do not have water molecules; and (2) how many explicit water
molecules are observed in the non-continuum solvent pockets defined in the membrane PBSA
calculation. The water probe (dprob) was scanned from 1.2 Å to 1.6 Å for three different
proteins: 1K4C, 5CFB, 5HCJ. The membrane setup has been optimized according to the values
given in Table II. For each protein, the samples of water molecules were taken from a 5ns
equilibrium MD simulation with all protein atoms restrained to the initial structure, which was
obtained from the last snapshot of the unconstrained normal MD, which is also the reference for
the water sampling run. The listed values are the averages of 100 snapshots evenly selected from
the 5ns MD simulation.
24
1K4C
5CFB
5HCJ
1.2
1.3
1.4
1.5
1.6
1.2
1.3
1.4
1.5
1.6
1.2
1.3
1.4
1.5
1.6
23
18
15
14
12
20
17
14
10
7
22
18
9
9
7
4
8
11
14
15
5
10
11
17
22
4
7
11
17
23
Figure 5 Discrepancy between implicit and explicit water simulations. The protein surface of
1K4C (blue) is overlaid with a bond representation and sampled water positions (yellow). Left: a
solvent region defined by the PBSA model but with no explicit water. Right: explicit water is
detected in a region where no solvent is defined in the PBSA model.
It
is
instructive
to
point
out
that
the
inconsistency
between
the
two
representations
may
be
due
to
the
setup
of
the
explicit
water
MD
simulation
and
also
to
the
limitations
of
MD
sampling
of
water
distributions.
First,
it
is
well
known
that
isolated
water
cavities
exist
in
the
protein
interior,
which
are
disconnected
from
the
bulk
water.
Unless
crystal
water
molecules
were
observed
and
retained
in
the
initial
setup
of
the
MD
simulations,
these
isolated
cavities
are
most
likely
modeled
as
water--‐free
due
to
the
default
closeness
tolerance
used
in
the
placement
of
explicit
water
molecules
when
building
the
topology
files.
This
issue
would
lead
to
the
type
(1)
mismatches
described
above.
Second,
although
protein
atoms
were
restrained
during
the
MD
simulations,
they
are
not
as
inflexible
as
frozen
hard
spheres
as
in
the
case
of
the
continuum
solvent
model
that
must
use
a
single
mean
structure
as
input.
Their
motions
allow
minor
structural
changes,
leading
to
the
opening
and
closing
of
buried
water
cavities.
If
the
mean
structure
25
happens
to
correspond
to
a
closed
form,
the
continuum
model
would
not
capture
the
water--‐accessible
cavity.
Finally,
the
protein
atom
cavity
radii
that
were
used
to
present
the
size
of
each
atom
were
chosen
to
be
best
for
energetics
and/or
stability
of
the
MD
simulations.
These
may
or
may
not
be
optimal
to
quantify
water
accessibility
in
the
protein
interior.
This
points
to
future
efforts
to
model
the
protein--‐water
interface
more
self--‐consistently
based
on
the
consistent
energy
model
as
defined
by
the
protein--‐water
force
field
used
in
both
explicit
and
implicit
simulations.
MMPBSA
calculations
of
binding
affinities
Finally,
as
an
illustration
of
our
new
continuum
membrane
model,
we
conducted
a
set
of
binding
free
energy
calculations
of
ten
different
ligands
independently
bound
to
a
potassium
channel
protein.
The
computed
binding
affinities
and
experimental
IC50
values
are
summarized
in
Table
IV.
The
correlation
analysis
between
computation
and
experiment
is
shown
in
Figure
6.
Both
the
classical
and
modern
nonpolar
solvent
models
(INP=1
and
INP=2
respectively)
were
tested,
and
the
correlations
for
these
two
methods
are
similar,
which
is
consistent
with
what
we
expected.
Overall
good
correlations
with
experiment
were
observed:
with
correlation
coefficients
of
0.79
for
INP=1
and
0.73
for
INP=2
(due
to
the
smaller
range
of
the
data).
Name
RTln(IC50)
mthick
Amitriptyline
Perhexline (PE0)
Perhexline (PE1)
-7.08
-7.23
-7.23
36.086
36.661
36.473
26
mcenter MMPBSA
(INP=1)
-38.23
-40.73
-41.96
-10.383
-1.620
-5.681
MMPBSA
(INP=2)
-9.84
-12.44
-13.63
Mizolastine
Loratadine
Domperidone
Terfenadine (TE0)
Terfenadine (TE1)
droperidol
Pimozide
Sertindole
Astemizole
-9.15
-9.58
-9.62
-9.82
-9.82
-10.61
-10.97
-11.13
-12.81
36.260
36.126
35.979
36.616
37.216
36.459
36.910
36.438
34.133
1.360
-0.969
0.707
-2.319
-0.611
3.308
0.190
-1.161
-0.043
-61.25
-52.48
-59.53
-69.67
-72.91
-59.04
-60.04
-62.91
-67.64
-24.59
-18.64
-20.69
-29.00
-30.56
-24.78
-20.09
-23.89
-26.21
Table IV: MMPBSA binding affinities (kcal/mol) in comparison with experiment (IC50). The
slab membrane geometry (the thickness and z-center in Å) compiled from the explicit solvent
MD simulation are also shown for each complex.
Figure 6: MMPBSA binding affinities compared with experimental measurements. Binding
affinities are in kcal/mol. Top: MMPBSA was computed with the classical nonpolar solvent
27
model (INP=1), the correlation coefficient is 0.79. Bottom: MMPBSA was computed with the
modern nonpolar solvent model (INP=2), the correlation coefficient is 0.73.
Timing
analysis
Finally,
we
conducted
a
timing
analysis
of
the
new
membrane
model.
Table
V
summarizes
the
average
CPU
times
over
100
frames
that
are
used
for
setting
up
the
dielectric
grids
with
or
without
the
membrane
model
in
the
MMPBSA
calculation
of
the
receptor.
We
can
see
the
average
time
for
the
surface
calculation
increases
by
more
than
four
times;
this
is
mainly
because
two
separate
SES
calls
are
made,
once
with
the
water
probe
and
once
with
the
membrane
probe.
Furthermore,
the
SES
calculation
with
the
much
larger
membrane
probe
is
behind
the
much
higher
cost
in
the
total
SES
time
due
to
the
longer
non--‐bonded
list
and
many
more
overlaps
among
larger
probe--‐augmented
atomic
volumes.
In
addition,
the
grid--‐
labeling
step
is
also
about
three
times
slower,
though
not
a
significant
portion
of
the
overall
CPU
cost.
Finally,
the
mapping
from
grid
labels
to
dielectric
constants
changes
little
due
to
the
virtually
linear
nature
of
the
algorithm
43.
Overall
the
PBSA
calculations
are
about
25%
slower
with
the
new
continuum
membrane
model
than
those
without
any
continuum
membrane
(i.e.
modeled
as
a
globular
protein)
for
the
tested
protein--‐ligand
binding
calculations.
Globular Protein Setup
3.76
1.61
0.21
Membrane Protein Setup
15.63
4.79
0.22
SES Calculations (s)
Grid Labeling (s)
EPS Mapping (s)
Table V: Average CPU times (in seconds) used in setting up the dielectric grid for 100 snapshots
in the MMPBSA calculation of the receptor. The membrane-free set up was run using
memopt=0, and the membrane setup was run using memopt=1 in Amber/PBSA.
28
Conclusions
We
have
proposed
a
new
continuum
membrane
model
for
Poisson--‐Boltzmann
calculations
of
biomolecules.
Major
improvements
from
the
standard
continuum
slab
model
are
the
following:
1)
explicit--‐solvent
MD
simulations
were
utilized
to
fine
tune
the
slab
model,
i.e.
its
exact
location
and
thickness,
to
best
reproduce
the
solvent
accessibility
and
the
water
accessible
channel;
2)
A
two--‐step,
two--‐probe
initial
grid
labeling
procedure
was
adopted
to
address
highly
different
accessibility
in
the
membrane
region
and
water
region;
and
3)
A
depth--‐first
search
algorithm
was
introduced
to
detect
the
water
pores/channels
automatically
based
on
the
initial
grid
labels.
This
procedure
follows
our
basic
algorithm
proposed
for
globular
proteins
and
does
not
add
significant
overhead
to
the
numerical
PB
calculations.
Given
the
revisions
proposed
above,
we
optimized
the
membrane
probe
value
and
the
slab
membrane
model
(i.e.
its
thickness)
to
best
reproduce
the
distributions
of
buried
water
molecules
in
the
membrane
region
as
sampled
in
explicit
water
MD
simulations.
Three
different
membrane
proteins
with
channels
were
utilized
in
this
optimization.
Our
analysis
showed
that
a
slab
membrane
model
using
the
mean
phosphate
atom
positions
as
the
membrane
boundary
and
the
smallest
membrane
probe
of
2.7
Å
caused
the
least
penetration
of
the
continuum
membrane
into
the
protein
interior.
Apparently,
the
solvent
accessibility
also
depends
on
how
the
continuum
water
is
modeled.
Thus,
we
used
a
water
distribution
map
from
an
explicit
water
MD
simulation
as
benchmark
data
to
evaluate
how
the
hard
sphere
SES
behaves
with
one
single
adjustable
parameter,
i.e.
the
water
solvent
probe.
The
same
three
membrane
channel
proteins
were
29
analyzed
to
address
this
question.
It
is
interesting
to
note
that
the
standard
value
for
the
water
solvent
probe
of
1.4
Å
is
very
reasonable,
which
gives
an
overall
minimum
number
of
inconsistencies
between
the
continuum
and
explicit
representations
of
water
distributions
in
the
membrane
channel
proteins,
at
least
in
the
water
accessible
pore/channel
regions
that
we
have
focused
on.
Finally,
we
conducted
a
set
of
binding
affinity
calculations
of
ten
different
ligands
independently
bound
to
a
potassium
channel
using
the
new
continuum
membrane
model.
Both
the
classical
and
modern
nonpolar
solvent
models
were
tested,
and
the
correlations
with
experiment
are
similar
with
both
models,
which
is
consistent
with
our
findings
in
globular
proteins.
Overall
good
correlations
with
experiment
were
observed,
with
correlation
coefficients
of
0.79
for
INP=1
and
0.73
for
INP=2.
Finally,
our
timing
analysis
showed
that
the
average
time
for
the
surface
calculation
increased
by
more
than
four
times.
The
grid--‐labeling
step
is
also
about
three
times
slower
even
though
it
is
not
a
significant
portion
of
the
overall
CPU
cost.
The
mapping
from
grid
labels
to
dielectric
constants
changed
little
due
to
the
virtually
linear
nature
of
the
algorithm.
Future
efforts
will
be
conducted
to
model
the
protein--‐water
interface
more
self--‐
consistently
based
on
the
consistent
energy
model
as
defined
using
the
protein--‐water
force
field
in
both
explicit
and
implicit
simulations.
This
work
is
supported
in
part
by
the
NIH
(GM093040
&
GM079383).
Acknowledgements
30
References
1.
Almen,
M.
S.;
Nordstrom,
K.
J.
V.;
Fredriksson,
R.;
Schioth,
H.
B.,
Mapping
the
human
membrane
proteome:
a
majority
of
the
human
membrane
proteins
can
be
classified
according
to
function
and
evolutionary
origin.
Bmc
Biol
2009,
7.
Arinaminpathy,
Y.;
Khurana,
E.;
Engelman,
D.
M.;
Gerstein,
M.
B.,
Computational
2.
analysis
of
membrane
proteins:
the
largest
class
of
drug
targets.
Drug
Discov
Today
2009,
14,
1130--‐1135.
Yildirim,
M.
A.;
Goh,
K.
I.;
Cusick,
M.
E.;
Barabasi,
A.
L.;
Vidal,
M.,
Drug--‐target
network.
3.
Nat
Biotechnol
2007,
25,
1119--‐1126.
4.
Overington,
J.
P.;
Al--‐Lazikani,
B.;
Hopkins,
A.
L.,
Opinion
--‐
How
many
drug
targets
are
there?
Nat
Rev
Drug
Discov
2006,
5,
993--‐996.
Davis,
M.
E.;
Mccammon,
J.
A.,
Electrostatics
in
Biomolecular
Structure
and
5.
Dynamics.
Chem
Rev
1990,
90,
509--‐521.
6.
Honig,
B.;
Sharp,
K.;
Yang,
A.
S.,
Macroscopic
Models
Of
Aqueous--‐Solutions
--‐
Biological
And
Chemical
Applications.
J.
Phys.
Chem.
1993,
97,
1101--‐1109.
Honig,
B.;
Nicholls,
A.,
Classical
Electrostatics
in
Biology
and
Chemistry.
Science
7.
1995,
268,
1144--‐1149.
8.
Beglov,
D.;
Roux,
B.,
Solvation
of
complex
molecules
in
a
polar
liquid:
An
integral
equation
theory.
Journal
of
Chemical
Physics
1996,
104,
8678--‐8689.
9.
Cramer,
C.
J.;
Truhlar,
D.
G.,
Implicit
solvation
models:
Equilibria,
structure,
spectra,
and
dynamics.
Chem
Rev
1999,
99,
2161--‐2200.
Bashford,
D.;
Case,
D.
A.,
Generalized
born
models
of
macromolecular
solvation
10.
effects.
Annual
Review
Of
Physical
Chemistry
2000,
51,
129--‐152.
11.
Baker,
N.
A.,
Improving
implicit
solvent
simulations:
a
Poisson--‐centric
view.
Curr.
Opin.
Struct.
Biol.
2005,
15,
137--‐143.
Chen,
J.
H.;
Im,
W.
P.;
Brooks,
C.
L.,
Balancing
solvation
and
intramolecular
12.
interactions:
Toward
a
consistent
generalized
born
force
field.
Journal
of
the
American
Chemical
Society
2006,
128,
3728--‐3736.
Feig,
M.;
Chocholousova,
J.;
Tanizaki,
S.,
Extending
the
horizon:
towards
the
efficient
13.
modeling
of
large
biomolecular
complexes
in
atomic
detail.
Theoretical
Chemistry
Accounts
2006,
116,
194--‐205.
14.
Koehl,
P.,
Electrostatics
calculations:
latest
methodological
advances.
Curr.
Opin.
Struct.
Biol.
2006,
16,
142--‐151.
Im,
W.;
Chen,
J.
H.;
Brooks,
C.
L.,
Peptide
and
protein
folding
and
conformational
15.
equilibria:
Theoretical
treatment
of
electrostatics
and
hydrogen
bonding
with
implicit
solvent
models.
Peptide
Solvation
and
H--‐Bonds
2006,
72,
173--‐+.
Lu,
B.
Z.;
Zhou,
Y.
C.;
Holst,
M.
J.;
McCammon,
J.
A.,
Recent
progress
in
numerical
16.
methods
for
the
Poisson--‐Boltzmann
equation
in
biophysical
applications.
Communications
in
Computational
Physics
2008,
3,
973--‐1009.
17.
Wang,
J.;
Tan,
C.
H.;
Tan,
Y.
H.;
Lu,
Q.;
Luo,
R.,
Poisson--‐Boltzmann
solvents
in
molecular
dynamics
simulations.
Communications
in
Computational
Physics
2008,
3,
1010--‐
1031.
Altman,
M.
D.;
Bardhan,
J.
P.;
White,
J.
K.;
Tidor,
B.,
Accurate
Solution
of
Multi--‐Region
18.
Continuum
Biomolecule
Electrostatic
Problems
Using
the
Linearized
Poisson--‐Boltzmann
31
Equation
with
Curved
Boundary
Elements.
Journal
of
Computational
Chemistry
2009,
30,
132--‐153.
19.
Cai,
Q.;
Wang,
J.;
Hsieh,
M.--‐J.;
Ye,
X.;
Luo,
R.
Chapter
Six
--‐
Poisson–Boltzmann
Implicit
Solvation
Models.
In
Annual
Reports
in
Computational
Chemistry,
Ralph,
A.
W.,
Ed.;
Elsevier:
2012;
Vol.
Volume
8,
pp
149--‐162.
Xiao,
L.;
Wang,
C.;
Luo,
R.,
Recent
progress
in
adapting
Poisson–Boltzmann
methods
20.
to
molecular
simulations.
Journal
of
Theoretical
and
Computational
Chemistry
2014,
13,
1430001.
Botello--‐Smith,
W.
M.;
Cai,
Q.;
Luo,
R.,
Biological
applications
of
classical
electrostatics
21.
methods.
Journal
of
Theoretical
and
Computational
Chemistry
2014,
13,
1440008.
22.
Forsten,
K.
E.;
Kozack,
R.
E.;
Lauffenburger,
D.
A.;
Subramaniam,
S.,
Numerical--‐
Solution
of
the
Nonlinear
Poisson--‐Boltzmann
Equation
for
a
Membrane--‐Electrolyte
System.
J.
Phys.
Chem.
1994,
98,
5580--‐5586.
Spassov,
V.
Z.,
Yan,
L.,
and
Szalma,
S.
,
Introducing
an
Implicit
Membrane
in
23.
Generalized
Born/Solvent
Accessibility
Continuum
Solvent
Models.
J
Phys.
Chem.
B
2002,
106.
Im,
W.,
Feigh,
M.,
and
Brooks
III,
C.
L.
,
An
Implicit
Membrane
Generalized
Born
24.
Theory
for
the
Study
of
Structure,
Stability,
and
Interactions
of
Membrane
Proteins.
Biophysical
Journal
2003,
85,
2900--‐18.
Tanizaki,
S.;
Feig,
M.,
A
generalized
Born
formalism
for
heterogeneous
dielectric
25.
environments:
Application
to
the
implicit
modeling
of
biological
membranes.
Journal
of
Chemical
Physics
2005,
122.
Tanizaki,
S.;
Feig,
M.,
Molecular
dynamics
simulations
of
large
integral
membrane
26.
proteins
with
an
implicit
membrane
model.
J
Phys
Chem
B
2006,
110,
548--‐556.
27.
Callenberg,
K.
M.;
Choudhary,
O.
P.;
de
Forest,
G.
L.;
Gohara,
D.
W.;
Baker,
N.
A.;
Grabe,
M.,
APBSmem:
A
Graphical
Interface
for
Electrostatic
Calculations
at
the
Membrane.
PloS
One
2010,
5.
Li,
L.;
Li,
C.;
Sarkar,
S.;
Zhang,
J.;
Witham,
S.;
Zhang,
Z.;
Wang,
L.;
Smith,
N.;
Petukh,
M.;
28.
Alexov,
E.,
DelPhi:
a
comprehensive
suite
for
DelPhi
software
and
associated
resources.
BMC
biophysics
2012,
5,
9.
29.
Warwicker,
J.;
Watson,
H.
C.,
Calculation
of
the
Electric--‐Potential
in
the
Active--‐Site
Cleft
Due
to
Alpha--‐Helix
Dipoles.
J
Mol
Biol
1982,
157,
671--‐679.
30.
Bashford,
D.;
Karplus,
M.,
Pkas
Of
Ionizable
Groups
In
Proteins
--‐
Atomic
Detail
From
A
Continuum
Electrostatic
Model.
Biochemistry
1990,
29,
10219--‐10225.
Jeancharles,
A.;
Nicholls,
A.;
Sharp,
K.;
Honig,
B.;
Tempczyk,
A.;
Hendrickson,
T.
F.;
31.
Still,
W.
C.,
Electrostatic
Contributions
To
Solvation
Energies
--‐
Comparison
Of
Free--‐Energy
Perturbation
And
Continuum
Calculations.
Journal
of
the
American
Chemical
Society
1991,
113,
1454--‐1455.
32.
Gilson,
M.
K.,
Theory
Of
Electrostatic
Interactions
In
Macromolecules.
Curr.
Opin.
Struct.
Biol.
1995,
5,
216--‐223.
Edinger,
S.
R.;
Cortis,
C.;
Shenkin,
P.
S.;
Friesner,
R.
A.,
Solvation
free
energies
of
33.
peptides:
Comparison
of
approximate
continuum
solvation
models
with
accurate
solution
of
the
Poisson--‐Boltzmann
equation.
J
Phys
Chem
B
1997,
101,
1190--‐1197.
34.
Tan,
C.;
Yang,
L.;
Luo,
R.,
How
well
does
Poisson--‐Boltzmann
implicit
solvent
agree
with
explicit
solvent?
A
quantitative
analysis.
Journal
of
Physical
Chemistry
B
2006,
110,
18680--‐18687.
32
35.
Cai,
Q.;
Wang,
J.;
Zhao,
H.--‐K.;
Luo,
R.,
On
removal
of
charge
singularity
in
Poisson--‐
Boltzmann
equation.
Journal
of
Chemical
Physics
2009,
130.
36.
Wang,
J.;
Cai,
Q.;
Li,
Z.--‐L.;
Zhao,
H.--‐K.;
Luo,
R.,
Achieving
energy
conservation
in
Poisson--‐Boltzmann
molecular
dynamics:
Accuracy
and
precision
with
finite--‐difference
algorithms.
Chemical
Physics
Letters
2009,
468,
112--‐118.
Ye,
X.;
Cai,
Q.;
Yang,
W.;
Luo,
R.,
Roles
of
Boundary
Conditions
in
DNA
Simulations:
37.
Analysis
of
Ion
Distributions
with
the
Finite--‐Difference
Poisson--‐Boltzmann
Method.
Biophysical
Journal
2009,
97,
554--‐562.
Ye,
X.;
Wang,
J.;
Luo,
R.,
A
Revised
Density
Function
for
Molecular
Surface
Calculation
38.
in
Continuum
Solvent
Models.
Journal
of
Chemical
Theory
and
Computation
2010,
6,
1157--‐
1169.
39.
Luo,
R.;
Moult,
J.;
Gilson,
M.
K.,
Dielectric
screening
treatment
of
electrostatic
solvation.
Journal
of
Physical
Chemistry
B
1997,
101,
11226--‐11236.
40.
Wang,
J.;
Tan,
C.;
Chanco,
E.;
Luo,
R.,
Quantitative
analysis
of
Poisson--‐Boltzmann
implicit
solvent
in
molecular
dynamics.
Physical
Chemistry
Chemical
Physics
2010,
12,
1194--‐1202.
41.
Hsieh,
M.
J.;
Luo,
R.,
Exploring
a
coarse--‐grained
distributive
strategy
for
finite--‐
difference
Poisson--‐Boltzmann
calculations.
Journal
of
Molecular
Modeling
2011,
17,
1985--‐
1996.
Cai,
Q.;
Ye,
X.;
Wang,
J.;
Luo,
R.,
On--‐the--‐Fly
Numerical
Surface
Integration
for
Finite--‐
42.
Difference
Poisson--‐Boltzmann
Methods.
Journal
of
Chemical
Theory
and
Computation
2011,
7,
3608--‐3619.
43.
Wang,
J.;
Cai,
Q.;
Xiang,
Y.;
Luo,
R.,
Reducing
Grid
Dependence
in
Finite--‐Difference
Poisson--‐Boltzmann
Calculations.
Journal
of
Chemical
Theory
and
Computation
2012,
8,
2741--‐2751.
44.
Liu,
X.;
Wang,
C.;
Wang,
J.;
Li,
Z.;
Zhao,
H.;
Luo,
R.,
Exploring
a
charge--‐central
strategy
in
the
solution
of
Poisson's
equation
for
biomolecular
applications.
Physical
Chemistry
Chemical
Physics
2013.
45.
Wang,
C.;
Wang,
J.;
Cai,
Q.;
Li,
Z.
L.;
Zhao,
H.;
Luo,
R.,
Exploring
High
Accuracy
Poisson--‐Boltzmann
Methods
for
Biomolecular
Simulations.
Computational
and
Theoretical
Chemistry
2013,
1024,
34--‐44.
46.
Xiao,
L.;
Cai,
Q.;
Ye,
X.;
Wang,
J.;
Luo,
R.,
Electrostatic
forces
in
the
Poisson--‐Boltzmann
systems.
The
Journal
of
Chemical
Physics
2013,
139,
094106.
Xiao,
L.;
Cai,
Q.;
Li,
Z.;
Zhao,
H.
K.;
Luo,
R.,
A
Multi--‐Scale
Method
for
Dynamics
47.
Simulation
in
Continuum
Solvent
I:
Finite--‐Difference
Algorithm
for
Navier--‐Stokes
Equation.
Chemical
Physics
Letters
2014,
submitted.
48.
Miertus,
S.;
Scrocco,
E.;
Tomasi,
J.,
Electrostatic
Interaction
of
a
Solute
with
a
Continuum
--‐
a
Direct
Utilization
of
Abinitio
Molecular
Potentials
for
the
Prevision
of
Solvent
Effects.
Chemical
Physics
1981,
55,
117--‐129.
49.
Hoshi,
H.;
Sakurai,
M.;
Inoue,
Y.;
Chujo,
R.,
Medium
Effects
on
the
Molecular
Electronic--‐Structure
.1.
the
Formulation
of
a
Theory
for
the
Estimation
of
a
Molecular
Electronic--‐Structure
Surrounded
by
an
Anisotropic
Medium.
Journal
of
Chemical
Physics
1987,
87,
1107--‐1115.
50.
Zauhar,
R.
J.;
Morgan,
R.
S.,
The
Rigorous
Computation
of
the
Molecular
Electric--‐
Potential.
Journal
of
Computational
Chemistry
1988,
9,
171--‐187.
33
51.
Rashin,
A.
A.,
Hydration
Phenomena,
Classical
Electrostatics,
and
the
Boundary
Element
Method.
J.
Phys.
Chem.
1990,
94,
1725--‐1733.
Yoon,
B.
J.;
Lenhoff,
A.
M.,
A
Boundary
Element
Method
for
Molecular
Electrostatics
52.
with
Electrolyte
Effects.
Journal
of
Computational
Chemistry
1990,
11,
1080--‐1086.
53.
Juffer,
A.
H.;
Botta,
E.
F.
F.;
Vankeulen,
B.
A.
M.;
Vanderploeg,
A.;
Berendsen,
H.
J.
C.,
The
Electric--‐Potential
Of
A
Macromolecule
In
A
Solvent
--‐
A
Fundamental
Approach.
Journal
of
Computational
Physics
1991,
97,
144--‐171.
54.
Zhou,
H.
X.,
Boundary--‐element
Solution
of
Macromolecular
Electorstatic--‐interaction
energy
Between
2
Proteins.
Biophysical
Journal
1993,
65,
955--‐963.
Bharadwaj,
R.;
Windemuth,
A.;
Sridharan,
S.;
Honig,
B.;
Nicholls,
A.,
The
Fast
55.
Multipole
Boundary--‐Element
Method
for
Molecular
Electrostatics
--‐
an
Optimal
Approach
for
Large
Systems.
Journal
of
Computational
Chemistry
1995,
16,
898--‐913.
Purisima,
E.
O.;
Nilar,
S.
H.,
A
Simple
yet
Accurate
Boundary--‐Element
Method
for
56.
Continuum
Dielectric
Calculations.
Journal
of
Computational
Chemistry
1995,
16,
681--‐689.
57.
Liang,
J.;
Subramaniam,
S.,
Computation
of
molecular
electrostatics
with
boundary
element
methods.
Biophysical
Journal
1997,
73,
1830--‐1841.
Vorobjev,
Y.
N.;
Scheraga,
H.
A.,
A
Fast
Adaptive
Multigrid
Boundary
Element
Method
58.
For
Macromolecular
Electrostatic
Computations
In
A
Solvent.
Journal
of
Computational
Chemistry
1997,
18,
569--‐583.
Cortis,
C.
M.;
Friesner,
R.
A.,
Numerical
Solution
Of
The
Poisson--‐Boltzmann
Equation
59.
Using
Tetrahedral
Finite--‐Element
Meshes.
Journal
of
Computational
Chemistry
1997,
18,
1591--‐1608.
Holst,
M.;
Baker,
N.;
Wang,
F.,
Adaptive
Multilevel
Finite
Element
Solution
Of
The
60.
Poisson--‐Boltzmann
Equation
I.
Algorithms
And
Examples.
Journal
of
Computational
Chemistry
2000,
21,
1319--‐1342.
61.
Baker,
N.;
Holst,
M.;
Wang,
F.,
Adaptive
multilevel
finite
element
solution
of
the
Poisson--‐Boltzmann
equation
II.
Refinement
at
solvent--‐accessible
surfaces
in
biomolecular
systems.
Journal
of
Computational
Chemistry
2000,
21,
1343--‐1352.
62.
Totrov,
M.;
Abagyan,
R.,
Rapid
Boundary
Element
Solvation
Electrostatics
Calculations
In
Folding
Simulations:
Successful
Folding
Of
A
23--‐Residue
Peptide.
Biopolymers
2001,
60,
124--‐133.
63.
Boschitsch,
A.
H.;
Fenley,
M.
O.;
Zhou,
H.
X.,
Fast
boundary
element
method
for
the
linear
Poisson--‐Boltzmann
equation.
J
Phys
Chem
B
2002,
106,
2741--‐2754.
Shestakov,
A.
I.;
Milovich,
J.
L.;
Noy,
A.,
Solution
Of
The
Nonlinear
Poisson--‐Boltzmann
64.
Equation
Using
Pseudo--‐Transient
Continuation
And
The
Finite
Element
Method.
Journal
of
Colloid
and
Interface
Science
2002,
247,
62--‐79.
65.
Lu,
B.
Z.;
Cheng,
X.
L.;
Huang,
J.
F.;
McCammon,
J.
A.,
Order
N
Algorithm
For
Computation
Of
Electrostatic
Interactions
In
Biomolecular
Systems.
Proceedings
of
the
National
Academy
of
Sciences
of
the
United
States
of
America
2006,
103,
19314--‐19319.
66.
Xie,
D.;
Zhou,
S.,
A
New
Minimization
Protocol
For
Solving
Nonlinear
Poisson–
Boltzmann
Mortar
Finite
Element
Equation.
BIT
Numerical
Mathematics
2007,
47,
853--‐871.
Chen,
L.;
Holst,
M.
J.;
Xu,
J.
C.,
The
Finite
Element
Approximation
Of
The
Nonlinear
67.
Poisson--‐Boltzmann
Equation.
Siam
Journal
on
Numerical
Analysis
2007,
45,
2298--‐2320.
68.
Lu,
B.;
Cheng,
X.;
Huang,
J.;
McCammon,
J.
A.,
An
Adaptive
Fast
Multipole
Boundary
Element
Method
for
Poisson--‐Boltzmann
Electrostatics.
Journal
of
Chemical
Theory
and
Computation
2009,
5,
1692--‐1699.
34
69.
Bajaj,
C.;
Chen,
S.--‐C.;
Rand,
A.,
An
Efficient
Higher--‐Order
Fast
Multipole
Boundary
Element
Solution
For
Poisson--‐Boltzmann--‐Based
Molecular
Electrostatics.
Siam
Journal
on
Scientific
Computing
2011,
33,
826--‐848.
70.
Bond,
S.
D.;
Chaudhry,
J.
H.;
Cyr,
E.
C.;
Olson,
L.
N.,
A
First--‐Order
System
Least--‐
Squares
Finite
Element
Method
for
the
Poisson--‐Boltzmann
Equation.
Journal
of
Computational
Chemistry
2010,
31,
1625--‐1635.
71.
Klapper,
I.;
Hagstrom,
R.;
Fine,
R.;
Sharp,
K.;
Honig,
B.,
Focusing
of
Electric
Fields
in
the
Active
Site
of
Copper--‐Zinc
Superoxide
Dismutase
Effects
of
Ionic
Strength
and
Amino
Acid
Modification.
Proteins
Structure
Function
and
Genetics
1986,
1,
47--‐59.
Davis,
M.
E.;
McCammon,
J.
A.,
Solving
The
Finite--‐Difference
Linearized
Poisson--‐
72.
Boltzmann
Equation
--‐
A
Comparison
Of
Relaxation
And
Conjugate--‐Gradient
Methods.
Journal
of
Computational
Chemistry
1989,
10,
386--‐391.
Nicholls,
A.;
Honig,
B.,
A
Rapid
Finite--‐Difference
Algorithm,
Utilizing
Successive
73.
over--‐Relaxation
to
Solve
the
Poisson--‐Boltzmann
Equation.
Journal
of
Computational
Chemistry
1991,
12,
435--‐445.
Luty,
B.
A.;
Davis,
M.
E.;
McCammon,
J.
A.,
Solving
the
Finite--‐Difference
Nonlinear
74.
Poisson--‐Boltzmann
Equation.
Journal
of
Computational
Chemistry
1992,
13,
1114--‐1118.
75.
Holst,
M.;
Saied,
F.,
Multigrid
Solution
of
the
Poisson--‐Boltzmann
Equation.
Journal
of
Computational
Chemistry
1993,
14,
105--‐113.
Holst,
M.
J.;
Saied,
F.,
Numerical--‐Solution
Of
The
Nonlinear
Poisson--‐Boltzmann
76.
Equation
--‐
Developing
More
Robust
And
Efficient
Methods.
Journal
of
Computational
Chemistry
1995,
16,
337--‐364.
Bashford,
D.,
An
Object--‐Oriented
Programming
Suite
for
Electrostatic
Effects
in
77.
Biological
Molecules.
Lecture
Notes
in
Computer
Science
1997,
1343,
233--‐240.
78.
Im,
W.;
Beglov,
D.;
Roux,
B.,
Continuum
Solvation
Model:
computation
of
electrostatic
forces
from
numerical
solutions
to
the
Poisson--‐Boltzmann
equation.
Comput.
Phys.
Commun.
1998,
111,
59--‐75.
Rocchia,
W.;
Alexov,
E.;
Honig,
B.,
Extending
The
Applicability
Of
The
Nonlinear
79.
Poisson--‐Boltzmann
Equation:
Multiple
Dielectric
Constants
And
Multivalent
Ions.
J
Phys
Chem
B
2001,
105,
6507--‐6514.
80.
Wang,
J.;
Luo,
R.,
Assessment
of
Linear
Finite--‐Difference
Poisson--‐Boltzmann
Solvers.
Journal
of
Computational
Chemistry
2010,
31,
1689--‐1698.
81.
Luo,
R.;
David,
L.;
Gilson,
M.
K.,
Accelerated
Poisson--‐Boltzmann
calculations
for
static
and
dynamic
systems.
Journal
of
Computational
Chemistry
2002,
23,
1244--‐1253.
Cai,
Q.;
Hsieh,
M.--‐J.;
Wang,
J.;
Luo,
R.,
Performance
of
Nonlinear
Finite--‐Difference
82.
Poisson--‐Boltzmann
Solvers.
Journal
of
Chemical
Theory
and
Computation
2010,
6,
203--‐211.
83.
Xiao,
L.;
Cai,
Q.;
Ye,
X.;
Wang,
J.;
Luo,
R.,
Electrostatic
forces
in
the
Poisson--‐Boltzmann
systems.
Journal
of
Chemical
Physics
2013,
139.
84.
Wang,
J.;
Cieplak,
P.;
Li,
J.;
Wang,
J.;
Cai,
Q.;
Hsieh,
M.;
Lei,
H.;
Luo,
R.;
Duan,
Y.,
Development
of
Polarizable
Models
for
Molecular
Mechanical
Calculations
II:
Induced
Dipole
Models
Significantly
Improve
Accuracy
of
Intermolecular
Interaction
Energies.
Journal
of
Physical
Chemistry
B
2011,
115,
3100--‐3111.
85.
Lu,
B.;
Holst,
M.
J.;
McCammon,
J.
A.;
Zhou,
Y.
C.,
Poisson--‐Nernst--‐Planck
Equations
For
Simulating
Biomolecular
Diffusion--‐Reaction
Processes
I:
Finite
Element
Solutions.
Journal
of
Computational
Physics
2010,
229,
6979--‐6994.
35
86.
Lu,
Q.;
Luo,
R.,
A
Poisson--‐Boltzmann
dynamics
method
with
nonperiodic
boundary
condition.
Journal
of
Chemical
Physics
2003,
119,
11035--‐11047.
Botello--‐Smith,
W.
M.;
Luo,
R.,
Applications
of
MMPBSA
to
Membrane
Proteins
I:
87.
Efficient
Numerical
Solutions
of
Periodic
Poisson–Boltzmann
Equation.
Journal
of
Chemical
Information
and
Modeling
2015,
55,
2187--‐2199.
Georgescu,
R.
E.;
Alexov,
E.
G.;
Gunner,
M.
R.,
Combining
conformational
flexibility
88.
and
continuum
electrostatics
for
calculating
pK(a)s
in
proteins.
Biophysical
Journal
2002,
83,
1731--‐1748.
Nielsen,
J.
E.;
McCammon,
J.
A.,
On
the
evaluation
and
optimization
of
protein
X--‐ray
89.
structures
for
pKa
calculations.
Protein
Sci
2003,
12,
313--‐326.
90.
Warwicker,
J.,
Improved
pK(a)
calculations
through
flexibility
based
sampling
of
a
water--‐dominated
interaction
scheme.
Protein
Sci
2004,
13,
2793--‐2805.
Tang,
C.
L.;
Alexov,
E.;
Pyle,
A.
M.;
Honig,
B.,
Calculation
of
pK(a)s
in
RNA:
On
the
91.
structural
origins
and
functional
roles
of
protonated
nucleotides.
J.
Mol.
Biol.
2007,
366,
1475--‐1496.
Swanson,
J.
M.
J.;
Henchman,
R.
H.;
McCammon,
J.
A.,
Revisiting
free
energy
92.
calculations:
A
theoretical
connection
to
MM/PBSA
and
direct
calculation
of
the
association
free
energy.
Biophysical
Journal
2004,
86,
67--‐74.
93.
Bertonati,
C.;
Honig,
B.;
Alexov,
E.,
Poisson--‐Boltzmann
calculations
of
nonspecific
salt
effects
on
protein--‐protein
binding
free
energies.
Biophysical
Journal
2007,
92,
1891--‐
1899.
94.
Brice,
A.
R.;
Dominy,
B.
N.,
Analyzing
the
Robustness
of
the
MM/PBSA
Free
Energy
Calculation
Method:
Application
to
DNA
Conformational
Transitions.
Journal
of
Computational
Chemistry
2011,
32,
1431--‐1440.
Luo,
R.;
Gilson,
H.
S.
R.;
Potter,
M.
J.;
Gilson,
M.
K.,
The
physical
basis
of
nucleic
acid
95.
base
stacking
in
water.
Biophysical
Journal
2001,
80,
140--‐148.
David,
L.;
Luo,
R.;
Head,
M.
S.;
Gilson,
M.
K.,
Computational
study
of
KNI--‐272,
a
potent
96.
inhibitor
of
HIV--‐1
protease:
On
the
mechanism
of
preorganization.
Journal
of
Physical
Chemistry
B
1999,
103,
1031--‐1044.
97.
Shivakumar,
D.;
Deng,
Y.
Q.;
Roux,
B.,
Computations
of
Absolute
Solvation
Free
Energies
of
Small
Molecules
Using
Explicit
and
Implicit
Solvent
Model.
J
Chem
Theory
Comput
2009,
5,
919--‐930.
98.
Nicholls,
A.;
Mobley,
D.
L.;
Guthrie,
J.
P.;
Chodera,
J.
D.;
Bayly,
C.
I.;
Cooper,
M.
D.;
Pande,
V.
S.,
Predicting
small--‐molecule
solvation
free
energies:
An
informal
blind
test
for
computational
chemistry.
Journal
of
Medicinal
Chemistry
2008,
51,
769--‐779.
99.
Korman,
T.
P.;
Tan,
Y.--‐H.;
Wong,
J.;
Luo,
R.;
Tsai,
S.--‐C.,
Inhibition
kinetics
and
emodin
cocrystal
structure
of
a
type
II
polyketide
ketoreductase.
Biochemistry
2008,
47,
1837--‐
1847.
100.
Luo,
R.;
Head,
M.
S.;
Given,
J.
A.;
Gilson,
M.
K.,
Nucleic
acid
base--‐pairing
and
N--‐
methylacetamide
self--‐association
in
chloroform:
affinity
and
conformation.
Biophysical
Chemistry
1999,
78,
183--‐193.
101.
Mardis,
K.
L.;
Luo,
R.;
Gilson,
M.
K.,
Interpreting
trends
in
the
binding
of
cyclic
ureas
to
HIV--‐1
protease.
J.
Mol.
Biol.
2001,
309,
507--‐517.
102.
Marshall,
S.
A.;
Vizcarra,
C.
L.;
Mayo,
S.
L.,
One--‐
and
two--‐body
decomposable
Poisson--‐
Boltzmann
methods
for
protein
design
calculations.
Protein
Sci
2005,
14,
1293--‐1304.
36
103.
Hsieh,
M.
J.;
Luo,
R.,
Physical
scoring
function
based
on
AMBER
force
field
and
Poisson--‐Boltzmann
implicit
solvent
for
protein
structure
prediction.
Proteins--‐Structure
Function
and
Bioinformatics
2004,
56,
475--‐486.
104.
Wen,
E.
Z.;
Luo,
R.,
Interplay
of
secondary
structures
and
side--‐chain
contacts
in
the
denatured
state
of
BBA1.
Journal
of
Chemical
Physics
2004,
121,
2412--‐2421.
105.
Wen,
E.
Z.;
Hsieh,
M.
J.;
Kollman,
P.
A.;
Luo,
R.,
Enhanced
ab
initio
protein
folding
simulations
in
Poisson--‐Boltzmann
molecular
dynamics
with
self--‐guiding
forces.
Journal
of
Molecular
Graphics
&
Modelling
2004,
22,
415--‐424.
106.
Lwin,
T.
Z.;
Luo,
R.,
Overcoming
entropic
barrier
with
coupled
sampling
at
dual
resolutions.
Journal
of
Chemical
Physics
2005,
123.
107.
Lwin,
T.
Z.;
Zhou,
R.
H.;
Luo,
R.,
Is
Poisson--‐Boltzmann
theory
insufficient
for
protein
folding
simulations?
Journal
of
Chemical
Physics
2006,
124.
108.
Lwin,
T.
Z.;
Luo,
R.,
Force
field
influences
in
beta--‐hairpin
folding
simulations.
Protein
Science
2006,
15,
2642--‐2655.
109.
Tan,
Y.--‐H.;
Luo,
R.,
Protein
stability
prediction:
A
Poisson--‐Boltzmann
approach.
Journal
of
Physical
Chemistry
B
2008,
112,
1875--‐1883.
110.
Tan,
Y.;
Luo,
R.,
Structural
and
functional
implications
of
p53
missense
cancer
mutations.
BMC
Biophysics
2009,
2,
5.
111.
Lu,
Q.;
Tan,
Y.--‐H.;
Luo,
R.,
Molecular
dynamics
simulations
of
p53
DNA--‐binding
domain.
Journal
of
Physical
Chemistry
B
2007,
111,
11538--‐11545.
112.
Wang,
J.;
Tan,
C.;
Chen,
H.--‐F.;
Luo,
R.,
All--‐Atom
Computer
Simulations
of
Amyloid
Fibrils
Disaggregation.
Biophysical
Journal
2008,
95,
5037--‐5047.
113.
Srinivasan,
J.;
Cheatham,
T.
E.;
Cieplak,
P.;
Kollman,
P.
A.;
Case,
D.
A.,
Continuum
solvent
studies
of
the
stability
of
DNA,
RNA,
and
phosphoramidate
--‐
DNA
helices.
Journal
of
the
American
Chemical
Society
1998,
120,
9401--‐9409.
114.
Kollman,
P.
A.;
Massova,
I.;
Reyes,
C.;
Kuhn,
B.;
Huo,
S.
H.;
Chong,
L.;
Lee,
M.;
Lee,
T.;
Duan,
Y.;
Wang,
W.;
Donini,
O.;
Cieplak,
P.;
Srinivasan,
J.;
Case,
D.
A.;
Cheatham,
T.
E.,
Calculating
structures
and
free
energies
of
complex
molecules:
Combining
molecular
mechanics
and
continuum
models.
Accounts
Chem
Res
2000,
33,
889--‐897.
115.
Gohlke,
H.;
Case,
D.
A.,
Converging
free
energy
estimates:
MM--‐PB(GB)SA
studies
on
the
protein--‐protein
complex
Ras--‐Raf.
Journal
of
Computational
Chemistry
2004,
25,
238--‐
250.
116.
Yang,
T.
Y.;
Wu,
J.
C.;
Yan,
C.
L.;
Wang,
Y.
F.;
Luo,
R.;
Gonzales,
M.
B.;
Dalby,
K.
N.;
Ren,
P.
Y.,
Virtual
screening
using
molecular
simulations.
Proteins--‐Structure
Function
and
Bioinformatics
2011,
79,
1940--‐1951.
117.
Miller,
B.
R.;
McGee,
T.
D.;
Swails,
J.
M.;
Homeyer,
N.;
Gohlke,
H.;
Roitberg,
A.
E.,
MMPBSA.py:
An
Efficient
Program
for
End--‐State
Free
Energy
Calculations.
Journal
of
Chemical
Theory
and
Computation
2012,
8,
3314--‐3321.
118.
Wang,
C.;
Nguyen,
P.
H.;
Pham,
K.;
Huynh,
D.;
Le,
T.--‐B.
N.;
Wang,
H.;
Ren,
P.;
Luo,
R.,
Calculating
protein–ligand
binding
affinities
with
MMPBSA:
Method
and
error
analysis.
Journal
of
Computational
Chemistry
2016,
n/a--‐n/a.
119.
Li,
C.;
Li,
L.;
Zhang,
J.;
Alexov,
E.,
Highly
efficient
and
exact
method
for
parallelization
of
grid‐based
algorithms
and
its
implementation
in
DelPhi.
Journal
of
Computational
Chemistry
2012,
33,
1960--‐1966.
37
120.
Jogini,
V.;
Roux,
B.,
Electrostatics
of
the
intracellular
vestibule
of
K+
channels.
J.
Mol.
Biol.
2005,
354,
272--‐288.
121.
Botello--‐Smith,
W.
M.;
Liu,
X.;
Cai,
Q.;
Li,
Z.;
Zhao,
H.;
Luo,
R.,
Numerical
Poisson--‐
Boltzmann
Model
for
Continuum
Membrane
Systems.
Chemical
Physics
Letters
2012.
122.
Botello--‐Smith,
W.
M.;
Luo,
R.,
Applications
of
MMPBSA
to
Membrane
Proteins
I:
Efficient
Numerical
Solutions
of
Periodic
Poisson--‐Boltzmann
Equation.
J
Chem
Inf
Model
2015,
55,
2187--‐2199.
123.
Homeyer,
N.;
Gohlke,
H.,
Extension
of
the
free
energy
workflow
FEW
towards
implicit
solvent/implicit
membrane
MM--‐PBSA
calculations.
Bba--‐Gen
Subjects
2015,
1850,
972--‐982.
124.
Greene,
D.;
Botello--‐Smith,
W.
M.;
Follmer,
A.;
Xiao,
L.;
Lambros,
E.;
Luo,
R.,
Modeling
Membrane
Protein–Ligand
Binding
Interactions:
The
Human
Purinergic
Platelet
Receptor.
The
Journal
of
Physical
Chemistry
B
2016,
120,
12293--‐12304.
125.
Callenberg,
K.
M.,
Choudhary,
O.
P.,
de
Forest,
G.
L.,
Gohara,
D.
W.,
Baker,
N.
A.,
and
Grabe,
M.
,
APBSmem:
A
Graphical
Interface
for
Electrostatic
Calculations
at
the
Membrane.
PLoS
One
2010,
5,
1--‐11.
126.
Richards,
F.
M.,
Areas,
Volumes,
Packing,
and
Protein--‐Structure.
Annual
Review
of
Biophysics
and
Bioengineering
1977,
6,
151--‐176.
127.
Connolly,
M.
L.,
ANALYTICAL
MOLECULAR--‐SURFACE
CALCULATION.
Journal
of
Applied
Crystallography
1983,
16,
548--‐558.
128.
Connolly,
M.
L.,
SOLVENT--‐ACCESSIBLE
SURFACES
OF
PROTEINS
AND
NUCLEIC--‐
ACIDS.
Science
1983,
221,
709--‐713.
129.
Gilson,
M.
K.;
Sharp,
K.
A.;
Honig,
B.
H.,
Calculating
the
Electrostatic
Potential
of
Molecules
in
Solution
--‐
Method
and
Error
Assessment.
Journal
of
Computational
Chemistry
1988,
9,
327--‐335.
130.
Rocchia,
W.;
Sridharan,
S.;
Nicholls,
A.;
Alexov,
E.;
Chiabrera,
A.;
Honig,
B.,
Rapid
grid--‐
based
construction
of
the
molecular
surface
and
the
use
of
induced
surface
charge
to
calculate
reaction
field
energies:
Applications
to
the
molecular
systems
and
geometric
objects.
Journal
of
Computational
Chemistry
2002,
23,
128--‐137.
131.
Swanson,
J.
M.
J.;
Mongan,
J.;
McCammon,
J.
A.,
Limitations
of
atom--‐centered
dielectric
functions
in
implicit
solvent
models.
J
Phys
Chem
B
2005,
109,
14769--‐14772.
132.
Tan,
C.
H.;
Yang,
L.
J.;
Luo,
R.,
How
well
does
Poisson--‐Boltzmann
implicit
solvent
agree
with
explicit
solvent?
A
quantitative
analysis.
J
Phys
Chem
B
2006,
110,
18680--‐18687.
133.
Wang,
J.;
Tan,
C.;
Chanco,
E.;
Luo,
R.,
Quantitative
analysis
of
Poisson--‐Boltzmann
implicit
solvent
in
molecular
dynamics
simulations.
Physical
Chemistry
Chemical
Physics
2010,
12,
In
press.
134.
Case,
D.
A.;
Betz,
R.
M.;
Botello--‐Smith,
W.;
Cerutti,
D.
S.;
T.E.
Cheatham,
I.;
Darden,
T.
A.;
Duke,
R.
E.;
Giese,
T.
J.;
Gohlke,
H.;
Goetz,
A.
W.;
Homeyer,
N.;
Izadi,
S.;
Janowski,
P.;
Kaus,
J.;
Kovalenko,
A.;
Lee,
T.
S.;
LeGrand,
S.;
Li,
P.;
Lin,
C.;
Luchko,
T.;
Luo,
R.;
Madej,
B.;
Mermelstein,
D.;
Merz,
K.
M.;
Monard,
G.;
Nguyen,
H.;
Nguyen,
H.
T.;
Omelyan,
I.;
Onufriev,
A.;
Roe,
D.
R.;
Roitberg,
A.;
Sagui,
C.;
Simmerling,
C.
L.;
Swails,
J.;
Walker,
R.
C.;
Wang,
J.;
Wolf,
R.
M.;
Wu,
X.;
Xiao,
L.;
York,
D.
M.;
Kollman,
P.
A.
Amber
16;
University
of
California,
San
Francisco:
2016.
135.
Case,
D.
A.;
Betz,
R.
M.;
Botello--‐Smith,
W.;
Cerutti,
D.
S.;
T.E.
Cheatham,
I.;
Darden,
T.
A.;
Duke,
R.
E.;
Giese,
T.
J.;
Gohlke,
H.;
Goetz,
A.
W.;
Homeyer,
N.;
Izadi,
S.;
Janowski,
P.;
Kaus,
J.;
Kovalenko,
A.;
Lee,
T.
S.;
LeGrand,
S.;
Li,
P.;
Lin,
C.;
Luchko,
T.;
Luo,
R.;
Madej,
B.;
38
Mermelstein,
D.;
Merz,
K.
M.;
Monard,
G.;
Nguyen,
H.;
Nguyen,
H.
T.;
Omelyan,
I.;
Onufriev,
A.;
Roe,
D.
R.;
Roitberg,
A.;
Sagui,
C.;
Simmerling,
C.
L.;
Swails,
J.;
Walker,
R.
C.;
Wang,
J.;
Wolf,
R.
M.;
Wu,
X.;
Xiao,
L.;
York,
D.
M.;
Kollman,
P.
A.
AmberTools
16;
University
of
California,
San
Francisco:
2016.
136.
Zhou,
Y.
F.;
Morais--‐Cabral,
J.
H.;
Kaufman,
A.;
MacKinnon,
R.,
Chemistry
of
ion
coordination
and
hydration
revealed
by
a
K+
channel--‐Fab
complex
at
2.0
angstrom
resolution.
Nature
2001,
414,
43--‐48.
137.
Huang,
X.;
Chen,
H.;
Michelsen,
K.;
Schneider,
S.;
Shaffer,
P.
L.,
Crystal
structure
of
human
glycine
receptor--‐alpha
3
bound
to
antagonist
strychnine.
Nature
2015,
526,
277--‐+.
138.
Laurent,
B.;
Murail,
S.;
Shahsavar,
A.;
Sauguet,
L.;
Delarue,
M.;
Baaden,
M.,
Sites
of
Anesthetic
Inhibitory
Action
on
a
Cationic
Ligand--‐Gated
Ion
Channel.
Structure
2016,
24,
595--‐605.
139.
Cavalli,
A.;
Poluzzi,
E.;
De
Ponti,
F.;
Recanatini,
M.,
Toward
a
pharmacophore
for
drugs
inducing
the
long
QT
syndrome:
Insights
from
a
CoMFA
study
of
HERG
K+
channel
blockers.
Journal
of
Medicinal
Chemistry
2002,
45,
3844--‐3853.
140.
Webb,
B.;
Sali,
A.
Comparative
Protein
Structure
Modeling
Using
MODELLER.
In
Current
Protocols
in
Bioinformatics;
John
Wiley
&
Sons,
Inc.:
2002.
141.
O'Donovan,
C.;
Martin,
M.
J.;
Gattiker,
A.;
Gasteiger,
E.;
Bairoch,
A.;
Apweiler,
R.,
High--‐
quality
protein
knowledge
resource:
SWISS--‐PROT
and
TrEMBL.
Briefings
in
bioinformatics
2002,
3,
275--‐84.
142.
Larkin,
M.
A.;
Blackshields,
G.;
Brown,
N.
P.;
Chenna,
R.;
McGettigan,
P.
A.;
McWilliam,
H.;
Valentin,
F.;
Wallace,
I.
M.;
Wilm,
A.;
Lopez,
R.;
Thompson,
J.
D.;
Gibson,
T.
J.;
Higgins,
D.
G.,
Clustal
W
and
Clustal
X
version
2.0.
Bioinformatics
(Oxford,
England)
2007,
23,
2947--‐8.
143.
Österberg,
F.;
Åqvist,
J.,
Exploring
blocker
binding
to
a
homology
model
of
the
open
hERG
K+
channel
using
docking
and
molecular
dynamics
methods.
FEBS
letters
2005,
579,
2939--‐2944.
144.
Du,
L.;
Li,
M.;
You,
Q.;
Xia,
L.,
A
novel
structure--‐based
virtual
screening
model
for
the
hERG
channel
blockers.
Biochemical
and
Biophysical
Research
Communications
2007,
355,
889--‐894.
145.
Stansfeld,
P.
J.;
Gedeck,
P.;
Gosling,
M.;
Cox,
B.;
Mitcheson,
J.
S.;
Sutcliffe,
M.
J.,
Drug
block
of
the
hERG
potassium
channel:
Insight
from
modeling.
Proteins:
Structure,
Function,
and
Bioinformatics
2007,
68,
568--‐580.
146.
Masetti,
M.;
Cavalli,
A.;
Recanatini,
M.,
Modeling
the
hERG
potassium
channel
in
a
phospholipid
bilayer:
Molecular
dynamics
and
drug
docking
studies.
J
Comput
Chem
2008,
29,
795--‐808.
147.
Jo,
S.;
Kim,
T.;
Im,
W.,
Automated
builder
and
database
of
protein/membrane
complexes
for
molecular
dynamics
simulations.
PLoS
One
2007,
2,
e880.
Jo,
S.;
Kim,
T.;
Iyer,
V.
G.;
Im,
W.,
CHARMM--‐GUI:
a
web--‐based
graphical
user
interface
148.
for
CHARMM.
J
Comput
Chem
2008,
29,
1859--‐65.
149.
Jo,
S.;
Lim,
J.
B.;
Klauda,
J.
B.;
Im,
W.,
CHARMM--‐GUI
Membrane
Builder
for
mixed
bilayers
and
its
application
to
yeast
membranes.
Biophys
J
2009,
97,
50--‐8.
150.
Wu,
E.
L.;
Cheng,
X.;
Jo,
S.;
Rui,
H.;
Song,
K.
C.;
Davila--‐Contreras,
E.
M.;
Qi,
Y.;
Lee,
J.;
Monje--‐Galvan,
V.;
Venable,
R.
M.;
Klauda,
J.
B.;
Im,
W.,
CHARMM--‐GUI
Membrane
Builder
toward
realistic
biological
membrane
simulations.
J
Comput
Chem
2014,
35,
1997--‐2004.
39
151.
Lee,
J.;
Cheng,
X.;
Swails,
J.
M.;
Yeom,
M.
S.;
Eastman,
P.
K.;
Lemkul,
J.
A.;
Wei,
S.;
Buckner,
J.;
Jeong,
J.
C.;
Qi,
Y.;
Jo,
S.;
Pande,
V.
S.;
Case,
D.
A.;
Brooks,
C.
L.,
3rd;
MacKerell,
A.
D.,
Jr.;
Klauda,
J.
B.;
Im,
W.,
CHARMM--‐GUI
Input
Generator
for
NAMD,
GROMACS,
AMBER,
OpenMM,
and
CHARMM/OpenMM
Simulations
Using
the
CHARMM36
Additive
Force
Field.
J
Chem
Theory
Comput
2016,
12,
405--‐13.
152.
Case,
D.
A.;
Cheatham,
T.
E.;
Darden,
T.;
Gohlke,
H.;
Luo,
R.;
Merz,
K.
M.;
Onufriev,
A.;
Simmerling,
C.;
Wang,
B.;
Woods,
R.
J.,
The
Amber
biomolecular
simulation
programs.
Journal
of
Computational
Chemistry
2005,
26,
1668--‐1688.
153.
Roe,
D.
R.;
Cheatham,
T.
E.,
PTRAJ
and
CPPTRAJ:
Software
for
Processing
and
Analysis
of
Molecular
Dynamics
Trajectory
Data.
J
Chem
Theory
Comput
2013,
9,
3084--‐95.
154.
Tan,
C.;
Tan,
Y.--‐H.;
Luo,
R.,
Implicit
nonpolar
solvent
models.
Journal
of
Physical
Chemistry
B
2007,
111,
12263--‐12274.
40
|
1112.2731 | 1 | 1112 | 2011-12-12T22:01:22 | Reactive Molecular Dynamics study on the first steps of DNA-damage by free hydroxyl radicals | [
"physics.bio-ph",
"physics.chem-ph",
"physics.comp-ph"
] | We employ a large scale molecular simulation based on bond-order ReaxFF to simulate the chemical reaction and study the damage to a large fragment of DNA-molecule in the solution by ionizing radiation. We illustrate that the randomly distributed clusters of diatomic OH-radicals that are primary products of megavoltage ionizing radiation in water-based systems are the main source of hydrogen-abstraction as well as formation of carbonyl- and hydroxyl-groups in the sugar-moiety that create holes in the sugar-rings. These holes grow up slowly between DNA-bases and DNA-backbone and the damage collectively propagate to DNA single and double strand break. | physics.bio-ph | physics | Reactive Molecular Dynamics study on the first steps of DNA-damage by free
hydroxyl radicals
Ramin M. Abolfath1,2, A. C. T. van Duin3, Thomas Brabec2
1School of Natural Sciences and Mathematics and Department of Materials Science,
University of Texas at Dallas, Richardson, TX 75080
2Physics Department, University of Ottawa,
Ottawa, ON, K1N 6N5, Canada
3Department of Mechanical and Nuclear Engineering,
Pennsylvania State University, PA 16802
(Dated: September 16, 2018)
1
1
0
2
c
e
D
2
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
1
3
7
2
.
2
1
1
1
:
v
i
X
r
a
We employ a large scale molecular simulation based on bond-order ReaxFF to simulate the chem-
ical reaction and study the damage to a large fragment of DNA-molecule in the solution by ionizing
radiation. We illustrate that the randomly distributed clusters of diatomic OH-radicals that are
primary products of megavoltage ionizing radiation in water-based systems are the main source of
hydrogen-abstraction as well as formation of carbonyl- and hydroxyl-groups in the sugar-moiety that
create holes in the sugar-rings. These holes grow up slowly between DNA-bases and DNA-backbone
and the damage collectively propagate to DNA single and double strand break.
PACS numbers: 87.56.-v,87.55.-x,87.64.Aa
It is known that megavoltage radiation (X/γ-rays, α-
particles, and heavy ions) ionizes the water molecule and
creates neutral free-radicals and aqueous electrons1 -- 9. In
particular OH-radicals with a very short life-time that is
reported to be within nano-seconds10, are major contrib-
utors to the single/double strand breaking of the DNA
molecules and the nucleotide-base damage, as 2/3 of envi-
ronment surrounding DNA molecules in the cell-nucleus
is composed of water molecules4. Various effects of the
ionizing radiation on biological systems that ranges from
the development of genetic aberrations, carcinogenesis to
aging, have attributed to the role of free radicals.
Computational modeling is a valuable tool in under-
standing the basic mechanisms that underlie DNA dam-
age. Great effort has been devoted to the statistical mod-
eling of the damage sites based on Monte-Carlo (MC)
sampling, using empirical reaction rates and radiation
scattering cross-sections4 -- 9. These models are limited to
MC sampling on a static structure of DNA or dynamical
models based on molecular-mechanics (MM) and empir-
ical force-fields (FF), e.g. AMBER/CHARMM FF, that
are developed for simulation of the non-reactive aspects
of bio-molecules.
The reactive aspects and the time evolution of the
multi-site DNA damages driven by cascade of chemical
reactions, that are beyond MM methods and empirical
FF, require the calculation of the potential energies on-
the-fly using first-principle quantum mechanical (QM)
models. Recently ab-initio simulations of the hydrogen
abstraction were developed11 -- 15. However realistic mod-
eling of DNA molecule with its environment requires ex-
tensive computer resources and is a major draw-back of
QM methods. The DFT calculation for hydrogen ab-
straction in vacuum is limited to the initial damage of a
single base11 -- 13 or single sugar-moiety15. Despite recent
advances in QM/MM methods that allows simulation of
larger molecules14 and inclusion of solvation15, the real
time simulation of DNA-damage remains still elusive.
To address the above considerations regarding to the
large scale modeling of DNA-damage, we have studied
the evolution of randomly distributed hydroxyl-radicals
in small pockets surrounding the DNA-molecule at room
temperature using molecular dynamics simulations where
the atomic interactions are described by the reactive force
field potential, ReaxFF17. ReaxFF is a general bond-
order dependent potential that provides accurate descrip-
tions of bond breaking and bond formation. Recent sim-
ulations on a number of hydrocarbon-oxygen systems17
and graphene-oxides18 showed that ReaxFF reliably pro-
vides energies, transition states, reaction pathways and
reactivity trends in agreement with QM calculations and
experiments.
To enable a reactive simulation of solvated DNA we
combined the recently developed ReaxFF reactive force
field parameters for peptide/water systems19 with the
ReaxFF description for organophosphates, as used pre-
viously by Zhu et al. 20 to investigate the active site of
RNA polymerase and by Quenneville et al.21 to study
dimethyl methylphosphonate (DMMP) reactions in sil-
ica. ReaxFF is a bond-order dependent empirical force
field method, which includes a polarizable charge func-
tion, enabling application to a wide range of materials
and accurate reproduction of reaction energies and barri-
ers17,22 -- 24. The ReaxFF peptide/phosphate descriptions
is fully transferable with the aqueous-phase ReaxFF de-
scriptions used by Raymand et al.25 to study water dis-
sociation reactions on zinc oxide surfaces and by Fogarty
et al.26 to study water reactions on silica surfaces, and
as such was tested against DFT-data describing proton
transfer reactions to solvated hydroxyl species.
To validate ReaxFF for hydroxyl ion/water interaction
and in order to describe the chemistry and physics of OH-
in the aqueous environment we compared the ReaxFF
parameters with a set of DFT-data. The DFT-data was
obtained at the X3LYP/6-311G** level30,31 and describes
two cases (1) binding energies of HO[H2O]n-clusters, with
n = 1 and (2) proton migration between HO-/H2O at
HO -- OH2 distances ranging from 2.4 to 3.4 A. While
ReaxFF overpredicts the OH-[H2O]n-binding energies for
n=1 and n=2, it gives an accurate description for the
n = 3 and n = 4 cases, which are the most relevant
for normal-density aqueous phases. ReaxFF displays ex-
cellent agreement for the proton migration barriers for
O -- O distances of 2.3, 2.48 (global minimum) and 2.8
A, which are the most relevant cases for normal-density
aqueous phases, but overpredicts the barriers for large
O -- O distances. We will provide a more elaborate de-
scription of the ReaxFF water/H3O+/OH− development
in a subsequent publication. The equilibration of the
system of DNA-water was performed in solvated 1276-
atom DNA-strand with 2500 water molecules at T=300K
for 30 ps. We found that during this time-frame the
DNA retained its overall helical configuration, indicating
that ReaxFF retains the overall structural integrity of
the DNA over such time-frames and that reactive events
observed during exposure to OH radicals can indeed be
associated with the radical reactivity.
We examined the stability of DNA surrounded
by water-molecules against ReaxFF by equilibrating
a solvated 1276-atom DNA-strand (with 2500 water
molecules) at room temperature for 30 ps. We found
that during this time-frame the DNA retained its overall
helical configuration, indicating that ReaxFF retains the
overall structural integrity of the DNA over such time-
frames and that reactive events observed during exposure
to OH radicals can indeed be associated with the radical
reactivity.
In this work we demonstrate that ReaxFF gives valu-
able insight on the details of the microscopic events
observed in agreement with the ab-initio CPMD and
QMMM CPMD-GROMACS calculation including: (1)
the DNA-backbone hydrogen abstraction and forma-
tion of water molecules (2) nucleotide-sugar-moiety bond
breaking and formation of carbonyl- and hydroxyl-groups
in sugar-moiety-rings (3) nucleotide-nucleotide hydrogen
bond disruption, (4) nucleotide structural damage, (5) ef-
ficiency of the radiation generated OH radicals in making
direct hydrogen-abstraction and (6) deactivation mecha-
nisms of OH-radicals within a cluster of radicals due to
formation of ozone-molecules and network of hydrogen
bonds.
Most significantly we illustrate that the collective dam-
age that breaks the bonds between nucleotide-bases and
DNA-backbone can be attributed to the formation of the
carbonyl- and hydroxyl-groups in the sugar-moiety rings.
The holes created by OH-radicals between the nucleotide-
bases and DNA-backbone grow up and evolve to large
holes that contain number of bases and a large segment
of DNA-backbone. It further propagates to the structural
base-damage, DNA single- and double-strand break.
2
FIG. 1:
(a) Initial structure of DNA surrounded by randomly
generated small pockets of OH-radicals shown in the circles.
(b) DNA structure after t = 0.24 ps. A large scale double
strand damage is clearly visible. Water molecules that are the
product of hydrogen-abstraction and two OH-OH bounded
by hydrogen bonds are shown in the circles. Carbon, oxygen,
nitrogen, phosphorus and hydrogen atoms are shown as green,
red, blue, gold and white, respectively.
FIG. 2:
(a) Small cluster of OH radicals close to DNA-
backbone is shown in circles. (b) One OH interacts with DNA-
backbone and the others combine in form of ozone molecule
H2O2. (c) Hydrogen abstraction from backbone is complete
and H2O molecule forms. Carbon, oxygen, nitrogen, phos-
phorus and hydrogen atoms are shown as green, red, blue,
gold and white, respectively.
Our model for the initial distribution of OH-radicals
surrounding DNA molecule is based on well established
description of water and ionizing-radiation interaction
in which a mega-voltage beam interacts with water
molecules through Compton effect and electron-positron
pair production and produce spurs and blobs identified
as small pockets of ion-pairs with typical diameter of 4-7
(a)
(b)
(a)
(b)
(c)
3
FIG. 3: A DFT calculation used for the confirmation of
the chemical pathways obtained by ReaxFF. Initial configura-
tion of OH-radical with deoxyribose sugar-moiety attached to
amino group (representing the nucleotide-base) shown in (a)
leads to formation of water molecule (b), hydrogen molecule
(c), hydroxyl group (d), and carbonyl group (e) as a func-
tion of initial
location of the OH-radical with respect to
sugar-moiety ring. The bond strain created by hydroxyl
and carbonyl groups lead to the creation of the hole in the
deoxyribose-ring. Carbon, oxygen, nitrogen and hydrogen
atoms are shown as green, red, blue and white, respectively.
nm that fit approximately 3-12 ion pairs1.
The information on the type of chemical reactions and
the time-evolution of the damage is collected via run-
ning the MD up to 30 ps where the rearrangement of the
atomic coordinates have been deduced from a dynamical
trajectory calculated by ReaxFF. These simulations per-
formed using periodic boundary conditions in a canonical
moles, pressure and temperature (NPT) ensemble with a
Nose-Hoover thermostat for temperature control and a
time step of 0.25 fs.
Fig. 1(a) shows the initial structure of the DNA-
molecule used in our simulation. The computational box
comprise of DNA taken from the protein data base33,
water molecules and randomly generated OH-radicals in
small clusters around the DNA molecule. The DNA
molecule and solvated system was first energy minimized
so as to eliminate any bad contact with water molecules
arising from solvation. Small clusters of OH-radicals
close to the DNA-backbone are shown within the circles.
Fig. 1(b) shows an early stage of the molecular-structure
at t = 0.24 ps, where the initial hydrogen-abstraction
from DNA-backbone takes place. As it is seen, one OH-
radical diffuses the distance (cid:96) ≈ 6A and removes H5(cid:48)
from the sugar-moiety, consistent with the experimental
data2,3 and CPMD calculation13. Another interesting
feature revealed in this simulation consists of the cor-
related states formed by OH-radicals that lead to their
passivation. For example in Fig. 1(b) we observe that
the two OH-radicals that do not participate in hydrogen-
FIG. 4:
(a) Initial configuration of an adenine in a double
strand DNA in the presence of OH-radicals (not shown in the
figure) and (b) the configuration of the distorted adenine with
a hole (formed by C-C broken bond) at 500fs. The numbers
represent the bond-length. The large separation of adenine
from the backbone and the stretched backbone that are indi-
cation of the base-damage and single-strand break are clearly
visible. For comparison an undamaged cytosine is shown be-
low the adenine. (c) The damaged cytosine, located one base
above the adenine shown in (a) at t = 3500 fs, illustrating
a collective damage to the base, deoxyribose-ring and DNA
strand. The opened gap in the deoxyribose pentagon-ring
corresponding to C4(cid:48) -O4(cid:48) missing bond is visible. It indicates
formation of stable carbonyl group. Carbon, oxygen, nitro-
gen, phosphorus and hydrogen atoms are shown as green, red,
blue, gold and white, respectively.
abstraction form hydrogen-bonds with smaller thermal
diffusion length than an isolated OH-radical. In a cluster
of OH-radicals, we find that not all of the OH-radicals
thermally diffuse and interact with DNA. Fig. 2 reveals
more details on H2O2 formation and their lower reac-
tivity with DNA-molecule. Consistent with our recent
QM/MM calculation in the solution15, we observe that
because of the hydrogen bond network forming between
OH-radicals and water molecules the time for hydrogen
abstraction is longer compare to similar simulation in
vacuum.
From ReaxFF-MD we realized that the initial damage
that evolves to the separation of the nucleotide-base from
DNA-backbone and DNA single strand break can be at-
tributed to the formation of the hydroxyl- and carbonyl-
groups in the sugar-rings in the backbone. The oxygen
from OH radical weaken the C-N bond that attaches
the sugar-moiety with the nucleotide-base. To investi-
gate such possibility and the consistency of ReaxFF with
DFT, we employed an ab-initio CPMD calculation in
vacuum as well as a QMMM CPMD-GROMACS that
allows adding realistic solution15. The results are shown
in Fig. 3. The molecular structure consists of deoxyri-
bose sugar-moiety attached to amino group in the pres-
(a)
(b)
(c)
(d)
(e)
(a)
(b)
(c)
4
Time-evolution of C4(cid:48) -O4(cid:48) bond distance, corre-
FIG. 5:
sponding to the opened gap in the pentagon-ring of deoxyri-
bose shown in Fig. 4(c). As a result of series of chemical
reactions due to OH-radicals, the C4(cid:48) -O4(cid:48) bond undergoes
strong fluctuations close to t ≈ 500 fs. It takes another 500 fs
that C4(cid:48) -O4(cid:48) bond breaks and the carbonyl-group forms. The
establishment of carbonyl-group and its dynamical stability
beyond t ≈ 1000 fs is clearly visible. The C4(cid:48) -O4(cid:48) distance in
open-ring fluctuates around dCO ≈ 1.8A due to thermal vi-
brations. For comparison, the time-evolution of C4(cid:48) -O4(cid:48) bond
distance corresponding to a deoxyribose far from OH-radicals
is shown (dashed lines). It shows an undamaged bond length
of dCO ≈ 1.4A, predicted by ReaxFF.
ence of OH-radical. Fig. 3 (d) and (e) show the result of
CPMD in vacuum on hydroxyl- and carbonyl-formation
and breaking of the bond in the sugar-ring that connects
the amino-group to the backbone.
Our CPMD consists of four stages of wave function
optimization, dynamical equilibration at T=300 K with
ionic temperature control that allows step-wise increase
of temperature, requenching of the wave function, and fi-
nally the microcanonical dynamics (constant energy en-
semble) as described in Ref.13. We have performed a
constant pressure and temperature MD simulation with
a reference temperature of 300K and a time step of 1fs.
Our CPMD wave-function optimization is implemented
in a plane-wave basis within local spin density approx-
imation (LSDA) with an energy cutoff of 75 Rydberg
(Ry), and with Becke27 exchange and Lee-Yang-Parr
(BLYP) gradient-corrected functional28. Norm conserv-
ing ultrasoft Vanderbilt pseudo-potentials were used for
oxygen, hydrogen, nitrogen and carbon. A cubic cell of
size (13A × 13A × 13A) is used together with Poisson
solver of Martyna and Tuckerman29 for the wave function
minimization in CPMD. For the QMMM calculation, the
sugar-moiety, guanine (as a representative of a nucleotide
base) and OH-radicals are in the QM part and the rest of
DNA fragment including phosphorous and its two oxygen
(O1P and O2P) and water molecules are in MM part. The
CPMD structural energy minimization and GROMACS
FIG. 6:
(a) Schematic diagram of the chemical pathway lead-
ing to single strand break observed in ReaxFF MD. DOXi and
DOXi+1 denote two neighboring deoxyribose pairs labeled by
the indices ith and i + 1th. Donation of oxygen from one OH-
radical lead to formation of free PO3 and two disconnected
DOX's. (b) P-O5(cid:48) DOXi and P-O5(cid:48) DOXi+1 bond distances as
a function of time shown by circles, selected from a site of
damage in which a cluster of OH-radicals are chemically ac-
tive. Simultaneous increase in bond distances is a signature
of the single strand break. For comparison the red dashed
line shows typical P-O5(cid:48) bond distance for a part of DNA
back-bone that is far from OH-radicals.
force-field are implemented for QM and MM parts. The
details of this calculation will be presented elsewhere15.
Fig. 4 focuses on a fragment of the DNA that the base-
sugar-moiety bond-breaking by OH-radicals is seen in
ReaxFF. The initial configuration of the base (an ade-
nine) is shown in Fig. 4(a). There are two OH-radicals
close to the sugar-moiety (not shown in the figure). Two
intermediate configurations are shown in Fig. 4(b) and
(c) corresponding to t = 500 fs and t = 3500 fs respec-
tively. Fig. 4(b) shows the damaged adenine and an un-
damaged cytosine that is one base below the adenine (far
from OH-radicals).
In Fig. 4(c) we show the damaged
adenine with another cytosine (one base above the ade-
nine). Both adenine and cytosine were initially close to
free radicals.
A substantial damage in adenine-groups into a ring
and chain is clearly observed. While some of this dis-
tortion may be related to secondary reactions related to
0200040006000800010000time (fs)11.21.41.61.82dCO (A)damaged siteundamaged sitePO5' -- DOXi+1 -- baseOOO5' -- DOXi -- baseHO-radicalOH-radicalH -- O5' -- DOXi+1 -- base -- OH -- O5' -- DOXi -- baseP -- OOO(a)0200040006000800010000time (fs)246810dPO(A)damaged site: PO5'-DOXidamaged site: PO5'-DOXi+1undamaged site(b)5
ab-initio CPMD and QMMM.
We now turn to DNA back-bone damage and obser-
vation of the single strand break (SSB) in ReaxFF MD.
We define dPO as the relative distance between P and
two neighboring deoxyribose-rings denoted by DOXi and
DOXi+1 as shown in Fig. 6(a). This figure summarizes a
pathway that leads to SSB observed in ReaxFF MD. Two
OH-radicals participate in the chemical reaction. One
OH donates an oxygen to phosphorous in the back-bone
that results to formation of PO3 and strong fluctuation
in the links between P and two neighboring DOX that
finally lead to disjointing two DOX's from back-bone as
shown in Fig. 6(b). Here solid lines marked by circles
show the time evolution of P-DOXi and P-DOXi+1 bond-
distances close to one of the clusters of OH-radicals. Note
that in our simulation SSB has been observed only in lo-
cations where OH-radicals actively participate in chem-
ical reactions. Clearly simultaneous stretch in P-DOXi
and P-DOXi+1 bonds is indication of concerted events
that lead to SSB. For comparison a typical time evolu-
tion of dPO is shown (the dashed-line) selected from an
undamaged site of DNA located far from OH-radicals.
As it is seen, ReaxFF predicts an average bond-length
of dPO ≈ 2A. Note that the cut-off, set for bond visu-
alization, chosen for carbon bonds with d = 1.5A. The
phosphate fragments seen in the figures are the artifact
of VMD visualization. The remaining H from the frag-
mented OH-radical either forms a water molecule by com-
bining with the other OH-radical [not shown in Fig. 6(a)]
or passivate the dangling bond of O5(cid:48)-DOXi and forming
a hydroxyl group. In the latter, the other OH disinte-
grates to H and O where H passivates the dangling bond
of O5(cid:48)-DOXi+1 and O oxidizes the base [see Fig. 6 (a)].
Fig. 7 reveals similar features in a larger scale. Here
the initial and final configurations corresponding to t = 2
ps are shown in Fig. 7(a) and (b). For comparison, a part
of DNA that is far from concentration of OH-radicals is
shown in Fig. 7(c). The damage and disconnectivity of
the bases-backbone is clearly visible in Fig. 7(b), where
the stability of the DNA in the absence of OH-radicals
is obvious from Fig. 7(c). The horizental/vertical blue
dash-lines show the base pair/stacking hydrogen bonds.
As shown in Fig. 7(b) the network of hydrogen bonds is
disrupted because of the damaged sites. In contrast, in
undamaged fragment, as shown in Fig. 7(c), the network
of hydrogen bonds are preserved by MD.
In conclusion, we have examined the MD simulation
based on ReaxFF on a large fragment of DNA molecule to
simulate the hydrogen abstraction by OH-radicals. The
simulation reveals various type of dehydrogenation from
DNA back-bone and DNA-base that evolve into the dis-
connectivity of the bases from the backbone due to car-
bonyl and hydroxyl formation in sugar-moiety by OH-
radicals. We further confirmed the results obtained by
ReaxFF-MD using an ab-initio CPMD calculation. Close
examination of damage with various length scales show
that the damage collectively disrupt the base-pair hy-
drogen bonds and lead to base-backbone bond breaking,
FIG. 7:
(a) Initial configuration of a DNA-fragment in the
presence of OH-radicals (not shown in the figure) and (b) the
configuration of that part at t =2000 fs with a pronounced
separated between bases and backbone. (c) For comparison
another part of the DNA that has no exposure with OH-
radicals is shown at t =2000 fs. Carbon, oxygen, nitrogen,
phosphorus and hydrogen atoms are shown as green, red, blue,
gold and white, respectively. The blue dash lines show the
inter-base and base-stacking hydrogen bounds.
the OH-damage, it is possible that the current ReaxFF
description overestimates the distortion rate. We aim
to replicate the distortion pathway at the DFT-level to
confirm the results obtained by ReaxFF and if neces-
sary will improve this aspect of the ReaxFF description
to make it consistent with DFT. We note that based on
our experimental results, such a substantial distortion is
qualitatively consistent with the absorption spectrum of
unirradiated and that receiving ionizing radiation where
the effect of irradiation can be visible in the blurring
of the absorbance peaks. Our quantitative comparison
between computational modeling and experimental data
obtained from photo-emission is on the way16. Finally
the observed damage in cytosine is mainly associated
with a dislocated hydrogen. As it is shown in the fig-
ure, both bases are moved away from sugar-moiety with
a distance that is approximately around 5 A, hence they
are completely disconnected from the backbone.
In Fig. 4(c) we observe an opened gap in deoxyribose-
ring due to C4(cid:48)-O4(cid:48) broken bond. The corresponding
time evolution of the bond-distance is shown in Fig. 5.
The bond undergoes a severe fluctuation in t ≈ 500 −
1000 fs due to occurrence of series of chemical reactions
by two OH-radicals initially located at opposite sides of
deoxyribose-ring including (a) hydrogen abstraction of
H3(cid:48) by OH radical incorporated by formation of water
molecule and (b) formation of hydroxyl group with C4(cid:48).
Finally beyond t ≈ 1000 fs the open-gap with dCO ≈ 1.8A
is established. These collection of events including base-
damage incorporated with formation of a hydroxyl- and
carbonyl-group is consistent with the picture deduced by
(a)
(b)
(c)
single-strand break and finally double strand break. The
present work that is in complement with previous calcu-
lation on smaller system in vacuum and solution using
DFT and QM/MM can be used as a computational plat-
form for energy scoring and the effects of radiation on
biological systems.
Authors would like to thank Dr. Kyeongjae Cho and
Dr. Yves Chabal for useful comments and discussion.
6
1 Hall, E. J. Radiobiology for the Radiologist, (Lippincott
Williams & Wilkins, Baltimore, Fifth Edition, 2000);
Becker, D.; Adhikary, A.; and Sevilla, M.; in Charge Mi-
gration in DNA, Ed. Chakraborty, T.; (Springer-Verlag,
Berlin, 2007).
2 Pogozelski, W. K.; Tullius, T. D.; Chem. Rev. 1998, 98,
1089.
3 Tullius, T. D. ; Greenbaum, J. A.; Curr. Opin. Chem. Biol.
2005, 9, 127.
4 Aydogan, B. ; Bolch, W. E.; Swarts, S. G.; Turner, J. E.;
and Marshall, D. T.; Radiat. Res. 2008, 169 223.
5 Friedland, W.; Jacob, P.; Paretzke, H. G.; Merzagora, M.;
and Ottolenghi, A.; Radiat. Environ. Biophys. 1999, 38,
39.
6 Terrisol, M.; and Beaudre, A.; Radiat. Prot. Dosimetry
1990, 31, 171.
17 van Duin, A. C. T.; Dasgupta, S.; Lorant, F.; Goddard,
W. A. Journal of Physical Chemistry A 2001, 105, 9396.
18 Bagri, A.; Mattevi, C.; Acik, M.; Chabal, Y. J.; Chhowalla,
M.; and Shenoy, V. B.; Nature Chem. 2010, 2, 581; Bagri,
A.; Grantab, R.; Medhekar, N. V.; and Shenoy, V. B.; J.
Phys. Chem. C 2010, 114, 12053.
19 Rahaman, O.; van Duin, A. C. T.; Goddard, W. A., III;
Doren, D. J. Journal of Physical Chemistry B 2011, ASAP.
20 Zhu, R.; Janetzko, F.; Zhang, Y.; van Duin, A. C. T.;
Goddard, W. A.; Salahub, D. R. Theoretical Chemistry
Accounts 2008, 120, 479.
21 Quenneville, J.; Taylor, R. S.; van Duin, A. C. T. Journal
of Physical Chemistry C 2010, 114, 18894.
22 LaBrosse, M. R.; Johnson, J. K.; van Duin, A. C. T. J.
Phys. Chem. A 2010, 114, 5855.
23 Chenoweth, K.; van Duin, A. C. T.; Goddard, W. A. Jour-
7 Wilson, W. E.; and Paretzke, H. G.; Radiat. Prot. Dosime-
nal of Physical Chemistry A 2008, 112, 1040.
try 1994, 52, 249.
8 Nikjoo, H.; and Charlton, D. E.; Calculation of range
and distributions of damage to DNA by high- and low-
LET radiations. In Radiation Damage to DNA: Struc-
ture/Function Relationships at Early Times (A. F. Fucia-
relli and J. D. Zimbrick, Eds.). Battelle Press, Columbus,
OH, 1995; Nikjoo, H.; O'Neill, P.; Terrissol, M.; and Good-
head, D.T.; Int. J. Radiat. Biol. 1994; 66, 453.
9 Semenenko, V. A.; and Stewart, R. D.; Radiat. Res. 2005,
164, 180; ibid. 2005, 164, 194.
10 Sies, H.; Europ. J. Biochem. 1993, 215, 213.
11 Mundy, C. J.; Colvin, M. E.; Quong, A. A.; J. Phys. Chem.
A. 2002, 106 10063; Wu, Y.; Mundy, C. J.; Colvin, M. E.;
Car, R.; J. Phys. Chem. A. 2004, 108 2922.
12 Close, D. M.; and Ohman, K. T.; J. Phys. Chem. A 2008,
112, 11207.
13 Abolfath, R. M.; J. Phys. Chem. B, 2009, 113, 6938; Abol-
fath, R. M.; and Brabec, T.; J. Compt. Chem. 2010, 31,
2601.
14 Gervasio, F. L.; Laio, A.; Iannuzzi, M.; and Parrinello, M.;
Chem. Eur. J. 2004, 10, 4846.
15 Abolfath, R. M.; Biswas, P. K.; Rajnarayanam, R.;
Brabec, T.; Kodym, R.; Papiez, L.; under prepration.
16 Abolfath, R. M.; Kodym, R.; unpublished.
24 Cheung, S.; Deng, W. Q.; van Duin, A. C. T.; Goddard,
W. A. Journal of Physical Chemistry A 2005, 109, 851.
25 Raymand, D.; van Duin, A. C. T.; Spangberg, D.; God-
dard, W. A.; Hermansson, K. Surf. Sci. 2010, 604, 741.
26 Fogarty, J. C.; Aktulga, H. M.; Grama, A. Y.; van Duin, A.
C. T.; Pandit, S. A. J. Chem. Phys. 2010, 132, 174704/1.
27 A. D. Becke, Phys. Rev. A 38, 3098 (1988).
28 C. Lee, W. Yang, and R. G. Parr, Phys. Rev. B 37, 785
(1988).
29 G. J. Martyna, M. E. Tuckerman, J. Chem. Phys. 110,
2810 (1999).
30 Bryantsev, V. S.; Diallo, M. S.; van Duin, A. C. T.; God-
dard, W. A. Journal of Chemical Theory and Computation
2009, 5, 1016.
31 Xu, X.; Goddard, W. A. Proc. Natl. Acad. Sci USA 2004,
101, 2673.
32 Hutter, J.; Ballone, P.; Bernasconi, M.; Focher, P.; Fois,
E.; Goedecker, S.; Parrinello, M.; Tuckerman, M. E.;
CPMD code, version 3.13, MPI fuer Festkoerperforschung,
Stuttgart IBM Zurich Research Laboratory, 1990-2008.
33 Zegar, I.S.; Setayesh, F.R.; DeCorte, B.L.; Harris, C.M.;
Harris, T.M.; Stone, M.P.; Biochemistry 1996, 35 4334;
http://www.pdb.org/pdb/
|
1809.10000 | 1 | 1809 | 2018-09-26T14:00:04 | Order parameter allows classification of planar graphs based on balanced fixed points in the Kuramoto model | [
"physics.bio-ph",
"cond-mat.stat-mech",
"nlin.AO"
] | Phase balanced states are a highly under-explored class of solutions of the Kuramoto model and other coupled oscillator models on networks. So far, coupled oscillator research focused on phase synchronized solutions. Yet, global constraints on oscillators may forbid synchronized state, rendering phase balanced states as the relevant stable state. If for example oscillators are driving the contractions of a fluid filled volume, conservation of fluid volume constraints oscillators to balanced states as characterized by a vanishing Kuramoto order parameter. It has previously been shown that stable, balanced patterns in the Kuramoto model exist on circulant graphs. However, which non-circulant graphs first of all allow for balanced states and what characterizes the balanced states is unknown. Here, we derive rules of how to build non-circulant, planar graphs allowing for balanced states from the simple cycle graph by adding loops or edges to it. We thereby identify different classes of small planar networks allowing for balanced states. Investigating the balanced states' characteristics, we find that the variance in basin stability scales linearly with the size of the graph for these networks. We introduce the balancing ratio as a new order parameter based on the basin stability approach to classify balanced states on networks and evaluate it analytically for a subset of the network classes. Our results offer an analytical description of non-circulant graphs supporting stable, balanced states and may thereby help to understand the topological requirements on oscillator networks under global constraints. | physics.bio-ph | physics | Order parameter allows classification of planar graphs based on balanced fixed points
in the Kuramoto model
Franz Kaiser and Karen Alim∗
Biological Physics and Morphogenesis Group, Max Planck Institute
for Dynamics and Self-Organization, 37077 Gottingen, Germany and
Institute for Nonlinear Dynamics, Faculty of Physics,
University of Gottingen, 37077 Gottingen, Germany
(Dated: September 27, 2018)
Phase balanced states are a highly under-explored class of solutions of the Kuramoto model and
other coupled oscillator models on networks. So far, coupled oscillator research focused on phase
synchronized solutions. Yet, global constraints on oscillators may forbid synchronized state, ren-
dering phase balanced states as the relevant stable state.
If for example oscillators are driving
the contractions of a fluid filled volume, conservation of fluid volume constraints oscillators to bal-
anced states as characterized by a vanishing Kuramoto order parameter. It has previously been
shown that stable, balanced patterns in the Kuramoto model exist on circulant graphs. However,
which non-circulant graphs first of all allow for balanced states and what characterizes the balanced
states is unknown. Here, we derive rules of how to build non-circulant, planar graphs allowing for
balanced states from the simple cycle graph by adding loops or edges to it. We thereby identify
different classes of small planar networks allowing for balanced states. Investigating the balanced
states' characteristics, we find that the variance in basin stability scales linearly with the size of
the graph for these networks. We introduce the balancing ratio as a new order parameter based on
the basin stability approach to classify balanced states on networks and evaluate it analytically for
a subset of the network classes. Our results offer an analytical description of non-circulant graphs
supporting stable, balanced states and may thereby help to understand the topological requirements
on oscillator networks under global constraints.
I.
INTRODUCTION
Weakly interacting systems are ubiquitous in nature.
They play a crucial role for oscillations and timing in bi-
ological systems [1]. Examples include circadian clocks,
cell metabolism, chemical oscillations or pacemaker cells.
Various models have been developed to describe such in-
teracting systems in terms of coupled oscillators [2, 3].
Here, the most well-studied model is the so-called Ku-
ramoto model due to its tractability [4 -- 6] and due to
the possibility to derive Kuramoto-like models on very
general grounds [6 -- 8].
Research regarding the Kuramoto model has focused
mainly on oscillator phase synchronization phenom-
ena [3, 4, 9]. Synchronization phenomena can be ob-
served in a broad variety of oscillator network topologies
and are comparably easy to study. Yet, little research has
addressed so-called balanced states, characterized by a
vanishing Kuramoto order parameter. These states have
mainly been of interest in the field of control theory [10 --
13], where they represent desired states to coordinate
autonomous vehicles. Yet, balanced states are of much
broader concern as they describe coupled oscillators with
global constraints that are preventing phase synchroniza-
tion. For example, such a global constraint is given by
conservation of fluid volume in a tubular network with
periodically contracting tube walls. A living example
are the contraction patters on the networks formed by
∗ [email protected]
the slime mold Physarum polycephalum [14]. Fascinat-
ingly, these living networks are dynamic in their topol-
ogy, which poses the question of how network topology
affects phase balanced states.
The only class of network topologies so far known to
support stable, balanced states in the Kuramoto model
are circulant graphs [9, 10]. However, many real-world
graphs are non-circulant, planar graphs. A dynamical
model leading to balanced states independent of the un-
derlying topology was introduced by Scardovi et al. [15]
as a modification to the Kuramoto model, but it intro-
duces an additional dynamical variable thus resulting in a
different model. Therefore, it is still unknown which non-
trivial planar network topologies support balanced states
Here, we construct planar graphs that support stable,
phase balanced states in the Kuramoto model. Starting
from the cycle graph as a prototype for graphs with bal-
anced stable fixed points in the Kuramoto model, we first
study the effect of loops added to such a graph and show
how these may be chosen to preserve balanced states.
A balanced state is only preserved if the loops repre-
sent the symmetries of the roots of unity found in many
balanced states. Subsequently, we probe how additional
edges change the balanced fixed points and identify the
possible changes leading to balanced topologies, which
are given by edges connecting two vertices of equal phase.
In particular, we demonstrate how these two building
blocks of balanced graphs may be combined to create a
large network of Kuramoto oscillators having stable, bal-
anced fixed points. For the graphs created this way, we
then show that the variance of basin stability in terms of
the winding numbers at stable fixed points scales linearly
8
1
0
2
p
e
S
6
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
0
0
0
0
1
.
9
0
8
1
:
v
i
X
r
a
with the size of the graph. Finally, we introduce the bal-
ancing ratio as a new measure that allows to compare
different graphs in terms of their balanced fixed points
using the basin stability approach. We are able to derive
analytical scaling laws for this parameter for the exam-
ined balanced graphs using the scaling calculated for the
variance. Thereby, we manage to capture the effect of
loops or additional edges on balanced fixed points in cy-
cle graphs in a quantitative manner. Our results show
that topologies other than circulant graphs may support
balanced fixed points and offer a new way of looking at
balanced topologies.
II. METHODS
A. Theoretical methods - The Kuramoto model on
complex networks
2
by CN for a cycle graph with N vertices. A cycle is a path
that starts and ends in the same vertex without passing
through any other vertex twice.
If the whole graph is
given by a single cycle, it is referred to as a cycle graph.
We are now ready to express the equation of motion
for the Kuramoto model on a graph G with N vertices
as
ϑ = ω − KB sin(cid:0)BT ϑ(cid:1),
where B is the graph's incidence matrix. Here, ϑ =
(ϑ1, ..., ϑN )T ∈ T N are the phase variables evolv-
ing dynamically on the N -torus T N over time, ω =
(ω1, ..., ωN )T ∈ RN is a vector of frequencies that is typ-
ically drawn from some frequency distribution g(ω) and
K is the coupling constant determining the strength of
the mutual interaction between different oscillators. We
assume this coupling constant to be identical for each in-
teraction. In general, one may shift to a rotating frame
such that 1T ω = N · (cid:104)ω(cid:105) = 0.
In order to study the Kuramoto model on networks, we
will first introduce some tools from graph theory [16, 17].
Consider a graph G(E, V ) consisting of N vertices with
vertex set V and edge set E. Then one may define its
adjacency matrix A ∈ NN×N by the components Aij
representing the number of edges starting in vertex i and
ending in vertex j. Furthermore, one can define the in-
cidence matrix B ∈ ZN×NE , where NE = E(G) is the
number of edges in G. It is defined by its components as
1,
−1,
0,
Bij =
if (j, i) ∈ E(G), i.e., ej → vi
if (i, j) ∈ E(G), i.e., ej ← vi
otherwise
,
where ej → vi denotes the fact that edge j is incident
in vertex i and vice versa. Furthermore, the Laplacian
matrix L ∈ ZN×N of a graph G is defined by
L = BBT .
A graph is called circulant, if its Laplacian matrix is a cir-
culant matrix. Circulant matrices are defined by the fact
that each row is a cyclic permutation of the previous one.
The eigenvalues λj and eigenvectors vj, j ∈ {1, ...N} of
the Laplacian matrix for circulant graphs take the simple
form of the roots of unity which reads [18],
N(cid:88)
lk(ρj)k,
λj =
vj =(cid:0)1, ρj, ρ2
k=1
j , ..., ρN−1
j
(cid:1)T
, j ∈ {0, ..., N − 1},
(1)
where ρj,N = ei2πj/N are the N th roots of unity, i.e. the
complex numbers such that (ρj,N )N = 1, ∀j.
In this publication, we focus exclusively on so-called
planar graphs. A graph is called planar if there exists
a drawing of the graph in which no two edges cross. A
particularly simple example of a circulant planar graph is
given by the cycle graph or simple cycle, which we refer to
Our main focus will be the Kuramoto model with zero-
frequency vector ω = 0. In this case, one may simply
assume a coupling constant equal to unity K = 1 which
can be achieved by rescaling time by the coupling con-
stant t(cid:48) = K · t. The new dynamics will then potentially
take shorter or longer times to reach a stable fixed point,
but will still follow the same dynamics. This leaves us
with the following simplified equation of motion
ϑ = −B sin(cid:0)BT ϑ(cid:1).
(2)
We will be interested in fixed points of this dynamics
which are characterized by a vanishing time derivative of
the phase variables ϑ = 0. Identifying such fixed points
which are balanced points at the same time is the main
goal of this work.
For coupled oscillator models, one can define a mea-
sure of synchrony of oscillators as already introduced by
Kuramoto [6] which is given by
p(ϑ) = R(t) · ei(cid:104)ϑ(t)(cid:105) =
eiϑj (t),
(3)
N(cid:88)
j=1
1
N
N
where (cid:104)ϑ(t)(cid:105) = 1
i=1 ϑi(t) denotes the average angle.
Typically, the length of this complex vector R(t) = p(ϑ)
is used as an order parameter for synchrony. This or-
der parameter assumes values between zero for balanced
states [10, 19], also termed incoherent [20], and one for
complete phase synchronization.
Based on this order parameter, we define the set of
balanced states B(N ) for a given number of oscillator N
by
ϑ ∈ TN(cid:12)(cid:12) N(cid:88)
j
.
B(N ) :=
eiϑj = 0
(4)
Consequently, a planar graph on which the Kuramoto
dynamics in Eq.(2) has a stable fixed point that is a bal-
anced state will be called a balanced graph in the follow-
ing. The interplay between the existence and stability of
(cid:80)N
fixed points in the Kuramoto model on the one hand -
which is highly dependent on the underlying topology -
and the state being balanced on the other hand - which
is a requirement unrelated to the topology - will be the
main interest of this work.
An observable that we will make use of in the follow-
ing sections is the winding number q. Assume that an
orientation was assigned to the planar graph underlying
the model resulting in an oriented graph Gσ and its cycle
basis, i.e. a basis of the graph's cycles space consisting
only of simple cycles [17], is given by BC = {C1, ..., CM}.
Then the winding number for some cycle Ck in the graph
reads
(cid:88)
(i,j)∈Ck
qk =
1
2π
∆ij,
(5)
where (i, j) are edges composing the cycle and ∆ij =
ϑi − ϑj is the phase difference along that edge taken
modulo 2π to project it onto the interval ∆ij ∈ (−π, π].
This winding number qk ∈ Z assumes integer values since
all phases along the path following the cycle need to be
uniquely defined and in order to retrieve the phase at the
starting point, the overall phase difference needs to total
to zero modulo 2π.
B. Stable fixed points of the Kuramoto dynamics -
the basin stability approach
In order to be able to study stable balanced fixed
points, we will classify a fixed point's stability based on
the concept of basin stability introduced in Ref. 21. It
is based on the fixed point's basin of attraction B, which
is the set of all initial conditions from which the system
converges to the fixed point. The basin's volume may
then be interpreted as the probability of returning to the
fixed point after a perturbation [21, 22]. In case of the
Kuramoto model, the overall phase space volume reads
V = [0, 2π]N . When estimating the basin of attraction
B(ϑ0) of some fixed point ϑ0 ∈ V , the fixed point's indi-
cator function needs to be evaluated in the whole phase
space. It reads for some ϑ(cid:48) ∈ V
(cid:40)
1B(ϑ0)(ϑ
(cid:48)) =
1, if ϑ(cid:48) ∈ B(ϑ0)
0, if ϑ(cid:48) (cid:54)∈ B(ϑ0)
.
The basin stability SB(ϑ0) of some fixed point ϑ0 ∈ TN
is then simply given by the integral over this indicator
function normalized by the overall phase space volume
SB(ϑ0) =
V 1B(ϑ0)(ϑ(cid:48)) dN ϑ(cid:48)
(2π)N
∈ [0, 1].
(6)
(cid:82)
A different notion typically used to find stable fixed
points in dynamical systems is linear stability analysis.
It is based on the eigenvalues of the dynamical system's
Jacobian matrix [23]. Based on this notion, one may cal-
culate the following sufficient criterion for a fixed point
3
to be linearly stable in the Kuramoto dynamics indepen-
dent of the underlying topology [10, 24]
cos(∆ϑe) > 0, ∀e ∈ E(G),
(7)
where ∆ϑe is the phase difference along some edge e in
the graph respecting its orientation.
C. The roots of unity as a special class of balanced
states
Little knowledge is available about the mathematical
structure of balanced states [11, 20]. Here, we will discuss
particular solutions for balanced states formed by the
roots of unity.
Consider a vector of phase variables forming the N th
roots of unity ϑ = (ϑ1, .., ϑN )T , where
ϑj =
2πj
N
, j ∈ {1, ..., N}.
, where (cid:80)
A well known property of the roots of unity ρj,N = eϑj
is the fact that they sum up to zero thus making the
corresponding angles balanced states, see section C in
the appendix for a proof. Since this is true for all roots
of unity, one may choose as well different subgroups of
different roots of unity ϑj = 2πj
k Nk = N ,
Nk
without changing the fact that they sum up to zero and
are thus solutions of balanced states [14]. Note that for
the particularly simple case where N = Nk, i.e. the case
of evenly distributed oscillators, this set of angles is also
referred to as splay states in the literature and plays an
important role in neuroscience [25, 26].
Importantly, the overall state remains balanced if one
adds potentially time-dependent angles αk(t) ∈ S1 to
each group of roots of unity summing up to zero sep-
arately
ϑ∗
j =
2πj
Nk
+ αk(t).
Note that it is possible to choose different αk for each
k roots of unity. Therefore, a subset BR(N )
group of N th
of the set of balanced states reads
(cid:110)
ϑ ∈ TN(cid:12)(cid:12)ϑj =
Nk = N, Nk ∈ N, αk(t) ∈ S1(cid:111) ⊂ B(N ).
(cid:88)
+ αk(t),
2πj
Nk
BR(N ) =
(8)
For small N < 5, all balanced states may be created this
way. The Nk may in general either be constructed using
the prime factors of the overall number of oscillators N ,
or any combination of prime numbers summing up to N
according to a theorem by Lam and Leung [27]. We will
show how these solutions may be realized on a graph-
level.
D. Computational methods - Monte Carlo method
and cycle flows allow to identify stable fixed points
In this work, we make use of two different methods to
analyze the Kuramoto model's fixed points numerically.
On the one hand, we use an algorithm presented
[28] that may be used to
recently by Manik et al.
find stable fixed points in the Kuramoto model.
It is
based on the fact that different solutions to the equa-
tion characterizing fixed points in the Kuramoto model
may only differ by constant flows around the underly-
ing graph's cycles [28 -- 30].
In addition to that, it al-
lows to establish a one-to-one correspondence between
fixed points and winding numbers in the case where all
phase differences between neighboring vertices are con-
tained within a π-interval centered around the origin
∆ij ∈ [−π/2, π/2], i, j ∈ E(G) [28, 30]. This also guar-
antees the fixed points to be stable due to the stability
criterion in Eq. 7. For many planar graphs, all stable
fixed points for the Kuramoto model with all-equal fre-
quencies will have phase differences contained in this in-
terval as established by a theorem due to Delabays et
al. [31].
On the other hand, we use Monte Carlo sampling to
determine a fixed point's basin stability. The phase space
in the Kuramoto model has a volume of V = [0, 2π]N for
N oscillators. Using the Monte Carlo method, M = 105
initial conditions were drawn at random from a uniform
distribution spanning V and the number of initial condi-
tions that converged to each of the stable fixed points was
counted to estimate the basin stability SB similar to the
approach in Ref. 32. Since this is a repeated Bernoulli
experiment, it carries the standard error [21],
σ(SB(ϑ0)) =(cid:112)(SB(ϑ0) · (1 − SB(ϑ0)))/M
where ϑ0 ∈ TN is the fixed point of interest and SB(ϑ0)
its basin stability. One can easily see that this error
reaches its maximal value σ(0.5) ≈ 2 · 10−3 at a basin
stability of SB(ϑ0) = 0.5 given the number of initial con-
ditions used here M = 105.
Integration of the Kuramoto equation of motion was
performed using a standard ODE solver contained in the
scipy package with stepsize dt = 0.5. Convergence to
a stable fixed point was ensured by checking that the
last change was small and calculating the Jacobian eigen-
values. When evaluating the balancing ratio, all fixed
point were counted as balanced if their order parameter
was smaller than a graph-specific threshold value. This
threshold value was calculated based on the order pa-
rameter at the smallest non-balanced fixed point for the
given graph, which varies a lot between different graphs.
Using these two approaches, we could identify all stable
fixed points that occupy a significant amount of the over-
all basin stability and may thus be found by the Monte
Carlo method or have all phase differences contained in
the above interval and are therefore identified by the al-
gorithm. In addition to that, we were able to find ana-
lytical approximations for the phase differences at stable
4
fixed points which we used to recheck in specific cases
that we did not miss stable balanced fixed points, which,
however, did not reveal any new balanced fixed points
not accounted for by either of the two methods.
III. RESULTS
A. Creating non-circulant balanced graphs from
cycle graphs
We will build non-circulant balanced graphs starting
with the simplest example of planar, circulant graphs,
the cycle graph and the 3-regular prism graph. The cy-
cle graph is denoted CN , where the index represents the
number of vertices in the graph N . A cycle is a path
that starts and ends in the same vertex consisting of dis-
tinct vertices V (CN ) = {v1, ..., vN} and edges of the form
E(CN ) = {(v1, v2), (v2, v3), ..., (vN , v1)} [17, p.8]. If the
whole graph is given by one cycle, we refer to the graph
as a cycle graph. For this graph and the version of the
Kuramoto equation of motion (2) used here, the phase
differences at fixed points of the dynamics may be calcu-
lated analytically. These phase differences at stable fixed
points read [13]
∆ϑc(qc) =
2πqc
N
, qc ∈ [−(cid:98)Nc/4(cid:99),(cid:98)Nc/4(cid:99)] ⊂ Z,
where qc are the winding numbers characterizing the
fixed point and the domain represents stable fixed points.
The 3-regular prism graphs is denoted YN , where the
index once again represents the number of vertices in
the graph. They may be created by connecting two cy-
cle graphs to each other at pairs of vertices. For these
graphs, phase differences at stable fixed points take the
same form as for the cycle graph with the crucial dif-
ference that the winding number qY characterizing sta-
ble fixed points is now smaller qY ∈ [−(cid:98)NY /8(cid:99),(cid:98)NY /8(cid:99)].
This is due to the fact that the number of stable fixed
points is limited by the number of vertices in the cen-
tral cycle which consists of half of the overall number of
vertices.
Importantly, the stable fixed points in these two
topologies are all balanced except for the synchronized
state. Starting from this observation, we will now con-
struct other, non-circulant balanced graphs.
1. Only a few building blocks may be used to create
balanced graphs from cycle graphs
As we have seen, all circulant graphs are balanced
graphs including the simple cycle. For this reason one
may try to construct other balanced graphs by modifying
the cycle graph. However, most modifications will make
the graphs non-balanced. In Fig. 1, two examples of bal-
anced and non-balanced graphs created from balanced
graphs are shown to highlight the difficulties of finding
5
FIG. 1. Balanced graphs may be constructed by making use
of the two different building blocks. (a) Balanced graph rep-
resenting the symmetries of the fourth roots of unity with
respect to the corresponding winding number q1 = +1 in the
spectral drawing.
(b) Graph consisting of two cycles con-
nected to each other but with additional loops added in such
a way, that they are symmetric with respect to each of the
big faces with winding numbers q2,1 = q2,2 = +1. (c) Two
cycle graphs connected at more than one edge do no longer
possess any stable balanced fixed points. Fixed point shown
with q3,1 = q3,2 = 1 is balanced if cycles are only connected at
a single edge. (d) A graph that is not symmetric with respect
to the winding number q4,1 = +1 of the big face. The stable
state shown is not balanced. Faces with no winding number
indicated carry winding number zero. All graphs shown here
have automorphism groups with basis elements being rotation
and reflection along horizontal axis.
appropriate modifications. The graphs shown in panels
c and d are non-balanced whereas the graphs in panels a
and b are although their overall symmetries in terms of
graph automorphisms are the same. They are given by
elementary rotation and reflection along the horizontal
symmetry axis.
Here, we will show how one may add edges to cycle
graphs, or add loops to them such that the resulting
graphs are non-circulant, but still balanced graphs. To
achieve this, we will make use of our knowledge about
a special class of balanced states constructed from the
roots of unity as defined in Eq. 8. In general, we found
two different ways of modifying cycle graphs that lead to
balanced graphs which we will discuss in the following
paragraph.
2. Loops added symmetrically to a cycle graph may result
in a balanced graph
An obvious way to modify cycle graphs is given by
adding loops consisting of several vertices to the graph.
Yet, most such operations will destroy the balancing
properties of fixed points. The only way to add loops to
the graph preserving this property is a symmetric addi-
tion of loops reflecting the symmetries of balanced states.
For example, if one was to add k loops consisting of m
FIG. 2.
(a) The cycle graph C18 to which k = 6 loops of
size m = 1 have been added with angular variables showing
a stable, phase balanced fixed point (colorbar). The fixed
point has winding number q = 1 in the central cycle.
(b)
Angular variables corresponding to the fixed point shown in
(a) plotted on the unit circle. The angles in each of the loops
are visible and form roots of unity themselves resulting in a
balanced fixed point.
vertices each to a cycle graph with a total of N vertices
in such a way that the resulting graph will be balanced,
we found that one may in general only do so by choos-
ing k as an integer divisor of N kN = n ∈ N. In this
case, one may choose the loops to be equidistant in terms
of the vertices between them, which corresponds to the
roots of unity in angular space. An example representing
the 6th roots of unity on the cycle graph C18 is shown in
Fig. 2. Panel (a) shows the cycle graph to which k = 6
loops consisting of m = 1 vertex each have been added
and the corresponding color code represents phase vari-
ables at this stable fixed point of the Kuramoto dynam-
ics. Panel (b) shows the corresponding angular variables
on the unit circle at this fixed point, which makes the one-
to-one correspondence between topology and the roots of
unity characterizing it obvious.
3. Adding edges to balanced graphs may preserve balanced
states
Suppose a balanced graph G was found along with the
corresponding stable balanced fixed point ϑ character-
ized by some winding number with absolute value larger
than one, q > 1, in one of the graph's cycles. If there
are two vertices that have exactly the same phase vari-
able ϑi = ϑj then one may add an edge to the graph
connecting these two vertices E(G(cid:48)) = E(G) ∪ {(i, j)}.
The corresponding fixed point in the new graph G(cid:48) will
still be a stable, balanced fixed point.
One may see that the new fixed point is still stable
as follows; suppose that J (ϑ) ∈ RN×N is the Jacobian
matrix characterizing the fixed point in the old system
before adding an edge with eigenvalues λk(J ) < 0, ∀k ∈
{1, ..., N}, which guarantees the fixed point's linear sta-
bility by using linear stability analysis [23]. If the edge
added to the graph reads (i, j) ∈ E(G(cid:48)), then the Jaco-
bian at the new fixed point J(cid:48)(ϑ(cid:48)) differs from the old
one by a matrix J(cid:48)(ϑ(cid:48)) = J (ϑ) + j which has the trivial
+1+1+1+1+1+10π2π3π22πabcdbalancednotbalanced+10π2π3π20π2π3π22πab6
created by adding cycles to the graph. In Fig. 3, exam-
ples of cycle graphs to which edges have been added with-
out destabilizing the corresponding balanced fixed points
are shown. Balanced fixed points in the cycle graph C20
characterized by winding numbers q = 2 (a), q = 3 (b)
and q = 5 (c) do not become unstable if edges are added
connecting vertices having the same phase variable (red
dotted lines). The winding number is distributed in cor-
respondence with the number of vertices to the new cycles
in the modified graphs.
Importantly, although all graphs shown here have non-
trivial automorphism groups, the two building blocks dis-
cussed here may also be combined to create graphs with
no trivial symmetry properties, see section A 1 a in the
appendix.
B. Towards a general classification of balanced
graphs using the balancing ratio
Now that we showed how a class of balanced graphs
may be constructed, we are going to quantify how bal-
anced they are. In general, there is little knowledge avail-
able about the structure of balanced fixed points in bal-
anced graphs apart from the fact that the Kuramoto or-
der parameter vanishes at such points. To be able to
compare different balanced graphs in an easily accessible
way, we introduce the balancing ratio as a new measure
based on the basin stability approach. We define this
as the fraction of basin stability occupied by all balanced
states in a graph G taken together
(cid:88)
SB(ϑ) ∈ [0, 1].
(9)
b(G) =
ϑ∈BG(N )
Here, BG(N ) is the set of balanced states consisting of N
angles, now referring exclusively to the stable balanced
states in the graph G as indicated by the subscript and
defined by Eq. 4. This measure may be understood as the
probability of ending up in a balanced state when starting
from a randomly chosen state. Importantly, most graphs
will have vanishing balancing ratio b(G) = 0 since they
are not balanced. In practice, the balancing ratio may
be calculated by performing repeated Monte Carlo ex-
periments and counting the number of times a balanced
state is reached.
1. Variance of basin stability scales linearly with number of
vertices for many balanced topologies
In order to calculate this balancing ratio of a circulant
graph for an arbitrary number of vertices in the graph,
we need to derive an expression for the basin stability of
the graph's fixed points. The basin stability SB(qc) in de-
pendence of the winding number qc for the fixed points
in the Kuramoto model on a cycle graph is known to
follow a Gaussian distribution SB(q) = 1√
2πσ2 e−q2/(2σ2)
FIG. 3. Additional edges (red, dotted lines) added to cycle
graphs may not change the overall stability properties of sta-
ble, balanced fixed points. (a) A stable, balanced fixed point
in a cycle graph (top) with winding number q1 = +2 (center
of graph) does not loose stability upon addition of an edge
connecting two vertices having the same angular value (bot-
tom). The overall winding number is distributed equally over
the two resulting graphs (bottom) q1,1 = q1,2 = +1 (cen-
ter of the graphs). (b) The same procedure may be applied
for higher order fixed points (top) by connecting again ver-
tices of equal phases (bottom). The overall winding number
q2 = +4 is again distributed on the different resulting graphs
with q2,1 = +1, q2,2 = +2 and q2,3 = +1 from left to right.
(c) The same mechanism can be applied starting from wind-
ing number q3 = +5 (center) and connecting vertices of same
phase. The last graph's automorphism group has only one ele-
ment which is the reflection along the horizontal axis, whereas
all other graphs have an additional rotational symmetry.
entries jkl = 1· (δkiδlj + δkjδli)− 1· (δkiδli + δkjδlj), i.e. a
symmetric matrix with non-zero entries only at the en-
tries characterizing the new edge. The entry unity is due
to the fact that the new edge has zero phase difference
such that cos(0) = 1. This matrix has only one non-zero
eigenvalue λ(j) = −2. Using Weyl's inequality for the
eigenvalues of symmetric matrices [33], one can see that
the Jacobian eigenvalues λ(cid:48)
k at the new fixed point fulfill
k ≤ λk + 0 due to the fact that zero is the maximal
λ(cid:48)
eigenvalue of the additional matrix j. Thus, all Jacobian
eigenvalues at the new fixed point are negative as well
and the fixed point is stable.
On the other hand, the fixed point stays balanced since
the new fixed point has exactly the same phase variables
ϑ(cid:48) = ϑ and the balancing condition is independent of
topology. An example is shown in Fig. 1,b, where an edge
was added to a cycle graph with loops. Using this pro-
cedure, we made use of the fact that one may subdivide
the overall balancing property into individual subgroups
summing to zero separately, see section II C.
If the underlying graph is a cycle graph, this procedure
may be applied if the winding number is an integer divi-
sor of N qN = m ∈ N. Then there are groups of q ver-
tices having the same angular value since q = N · ϕ/(2π),
where ϕ = ϑi − ϑi−1, ∀i, implies that ϑi = ϑi+m mod 2π,
i.e. each angular value is repeated at every m-th posi-
tion. It is important to note that although the addition
of edges as described above does not change the stabil-
ity property of the corresponding fixed point it will affect
the basin stability of a fixed point as new fixed points are
+2+4+5+1+1+2+1+1+2+2+10π2π3π22π⇓⇓⇓abc7
if plotted against the winding numbers of the two cycles
q1 and q2.
If the two cycles have the same number of
vertices n = N/2 + 1, the Gaussian distribution is well-
described by a single standard deviation σ1 = σ2 = σ.
This variance shows the same scaling as for the single
cycle if plotted against the number of vertices in each of
the cycles n. In terms of the number of vertices n in a
cycle, it reads
(n) = (3.2 · n − 1.1) · 10−2.
σ2
Gc
This scaling is shown in terms of the overall number of
vertices N in Fig. 4 a (blue).
Finally, this scaling is also valid for graphs to which
loops have been added according to the scheme described
in section III A 2. The variance of the Gaussian scaling
observed in terms of the winding number was evaluated
for the winding number in the central loop. In terms of
the number of vertices in this central loop N∗ = N−k·m,
where N is the overall number of vertices in the graph,
k is the number of loops added and m is the number of
vertices in each loop, the overall variance shows a similar
scaling compared to the cycle graph
(N∗) = (3.1 · N∗ − 4.9) · 10−2.
σ2
Gl
This scaling was obtained using a linear fit on the Monte
Carlo results of 18 different graphs with k ∈ {2, 3, 5}
loops and different number of vertices.
It is shown in
Fig. 4 a (purple).
2. Balancing ratio for circulant graphs shows a simple
square root scaling
Inspired by the numerical results, we will now proceed
to show how one may calculate the balancing ratio in
circulant graphs. Importantly, all states except for the
synchronized state with winding number qc = 0 are bal-
anced for circulant graphs. Therefore, the balancing ra-
tio is calculated by summing over all stable states' basin
stability except for the synchronized state which is the
√
peak of the Gaussian distribution. This peak is of height
2πσCN (N ))−1 for a normalized Gaussian distribution
(
where σCN (N ) is the Gaussian's standard deviation. Us-
ing the linear scaling for the variance σ2
(N ), we get
the following expression for the balancing ratio of a cycle
graph CN in dependence of the overall number of vertices
in the graph
CN
SB(q) = 1 −(cid:16)√
(cid:88)
(cid:17)−1
≈ 1 −(cid:16)(cid:112)2π(3.2 · 10−2N )
q(cid:54)=0
.
(cid:17)−1
b(CN ) =
2πσCN (N )
(10)
Here, we neglected the y-intercept due to its negligible
effect on the result. This approximation is shown for
different numbers of vertices N in Fig. 6 (purple).
It
coincides very well with the results from Monte Carlo
sampling.
(a) Scaling of the variance σ2 characterizing the
FIG. 4.
Gaussian distribution describing the basin stability with the
number of vertices N in different graphs. Linear fits (dotted
lines) and data points from Monte Carlo sampling (mark-
ers) are shown for cycle graphs CN (dark purple dots), cycle
graphs with loops Gl (light purple squares), a cycle graph
with a central edge added Gc (blue diamonds) and prism
graphs YN (green stars). Fit parameters for linear relationship
σ2(N ) = m · N + b are mCN = 3.2 · 10−2, bCN = −1.7 · 10−3,
mGl = 3.1··10−2, bGl = −4.9·10−2, mGc = 1.6·N·10−2, bGc =
−1.1· 10−3 and mYN = 1.3· 10−2, bYN = −1.0· 10−2 from top
to bottom. Note that for Gl, fit was performed in terms of the
number of vertices in the central cycle. (b) Scaling of share
of basin stability occupied by the stable fixed points with all
winding numbers in outer loops equal to zero AG(k) in depen-
dence of the number of loops k added to a cycle graph. Data
points from Monte Carlo sampling represented by dots and
analytical approximation as dotted line are shown. Errors in
both figures take a maximal value of σ(SB(ϑ)) = 2 · 10−3 for
each dot and are thus too small to be visible in the figure.
and different explanations for this scaling have been sug-
gested [22, 34]. As we found here, the variance σ2
(N )
of this Gaussian distribution scales to good approxima-
tion linearly with the number of vertices N as shown in
Fig. 4a (dark purple line), i.e.,
CN
(N ) = 3.2 · 10−2 · N − 1.7 · 10−3.
σ2
CN
This fit was calculated by fitting a linear function to
the results obtained using Monte Carlo sampling and
calculating the variance in the winding number distri-
bution [35]. For moderate values of N , the scaling is
(N ) = 3.2 · 10−2 · N effectively since the y-intercept
σ2
is negligible.
CN
In addition to the linear scaling observed for the vari-
ance in case of the cycle graph, we found this scaling to
hold for other graphs as well. For the prism graph YN ,
(N ) = (1.3 · N − 1.0) · 10−2
the variance scales as σ2
if fitted linearly to the data. This result coincides very
well with the data points obtained from simulation, see
Fig. 4,a, green stars.
YN
Another example for which we found the linear scaling
to hold is the graph created by adding a single edge to
a cycle graph thus creating two cycles of the same size
connected at a single edge as described in section III A 3
and shown in Fig. 3 a, bottom. In this case, the basin
stability follows a two-dimensional Gaussian scaling
SB(q1, q2) =
1
2πσ1σ2
e−1/2((q1/σ1)2+(q2/σ2)2),
20406080100N01234σ2linearfitMCCNMCGlMCGcMCYN0102030k0.70.80.91.0AG(k)predictionMCsimulationsab8
Along the same lines, one may calculate a similar scal-
ing law for the prism graph YN . Again plugging in the
linear scaling obtained for the variance, the balancing
ratio for this graph is calculated to be
SB(q) = 1 −(cid:16)√
(cid:17)−1
(cid:88)
≈ 1 −(cid:16)(cid:112)2π(1.3 · N − 1.0) · 10−2)
(cid:17)−1
2πσYN (N )
q(cid:54)=0
(11)
.
b(YN ) =
Here, q is the winding number in the central cycle of
the prism graph. This approximation is shown in Fig. 6
(dark green line) again coinciding with the results ob-
tained from Monte Carlo sampling.
=
=
making use of the fact that the basin stability in terms of
the winding numbers follows a two-dimensional Gaussian
distribution and the scaling of this Gaussian's variance
in terms of the number of vertices n is again linear. This
allows us to evaluate the balancing ratio for the graph
Gc
b(Gc) =
(cid:88)
(cid:88)
(cid:88)
q∈Dc(q)
1
q∈[1,(cid:98)n/4(cid:99)]
q∈[1,(cid:98)n/4(cid:99)]
1
1 +q2
2 )/(2σ2)δq1,q2 − 1
2πσ2
2πσ2 e−(q2
πσ2 e−q2/σ2
π(3.2 · n − 1.1) · 10−2 e−q2/(3.2·n−1.1)·10−2
1
.
(12)
3. Balancing ratio depends on newly created cycles when
adding edges to cycle graphs
Having solved balancing ratios for simple circulant
graphs, we now consider the balanced graph obtained
by adding a single edge to a cycle graph CN such that
each of the newly created cycles has the same number of
vertices n = N/2 + 1. We will refer to this graph by Gc.
This graph is the most simple balanced graph one may
obtain using the above building block of adding edges to
balanced graphs and thus provides a natural extension
of cycle graphs. Exemplarily, such a graph is shown in
Fig. 3 a, bottom. Making use of the results obtained for
the cycle graph, we will show how one may calculate the
balancing ratio for this graph.
Fixed points in this graph are characterized by the
winding numbers in the two cycles q = (q1, q2)T . In this
case, the only balanced states that we found were states
where the two cycles have the same winding number,
i.e. opposite flows at the shared edge, which are precisely
the fixed points described in section III A 3 and in one-
to-one correspondence with the balanced fixed points in
the cycle graph without additional edge. Here, we assume
the cycles and edges to be both oriented counterclockwise
and the edge shared between the cycles to be oriented
along the first cycle characterized by q1. We define the
set characterizing all states where the two cycles have
identical winding number by
Qc = {q ∈ Dc(q)(cid:12)(cid:12)q1 = q2, q1 (cid:54)= 0}.
4(cid:99)(cid:3)2 ⊂ Z2 is the domain where
Here, Dc(q) = (cid:2)−(cid:98) n
4(cid:99),(cid:98) n
the fixed points are guaranteed to be stable, but not all
combinations of winding numbers in this set may nec-
essarily be realized in a stable fixed point. Again, the
synchronized state is not balanced for this topology and
thus excluded from this set. Using this definition, the
balancing ratio may be calculated as
(cid:88)
(cid:88)
b(Gc) =
SB(ϑ) =
SB(q)δq1,q2−SB((0, 0)T ),
ϑ∈BGc (N )
q∈Dc(q)
This result is shown in Fig. 6, (dark blue) and agrees very
well with the result obtained from Monte Carlo sampling.
Comparing this result to the one obtained for the cycle
graph CN in Eq. 10, one may notice that it is always
smaller by a prefactor of (2π)−1/2. In addition to that,
the variance σ2(N ) shows a more moderate scaling with
the number of vertices for the given graph and the sum
scales with σ−2 ∝ N−1 in the prefactor compared to the
σ−1 ∝ N−1/2 scaling observed for the cycle graph. This
explains the different scalings observed in Fig. 6.
4. Balancing ratio for cycle graph with loops depends on
graph symmetries
In this section, we will show how a similar result may
be obtained for cycle graphs to which loops have been
added. Consider a cycle graph to which k loops have
been added by adding m = 3 vertices to the graph for
each of the loops and connecting them to a single edge
of the graph, thus creating loops consisting of m + 2 = 5
vertices, see e.g. Fig. 5 b, for an example with two loops.
Each such loop results potentially in the creation of two
new stable fixed points characterized by the two non-
zero loop winding numbers ql ∈ {1,−1}. For this reason,
stable fixed points with all phase differences contained
in the interval [−π/2, π/2] centered around the origin
may be uniquely characterized using the winding vector
q = (q1, ql,1, ..., ql,k)T , where q1 is the winding number in
the central loop and the ql,k are the k winding numbers
characterizing the loops.
In order to be able to calculate the balancing ratio for
such graphs, we will first study the effects of loops on
the overall basins stability occupied by the newly cre-
ated fixed points. The basin stability occupied by such
a stable state in which exactly one loop has a non-zero
winding number stays roughly constant over all numbers
of vertices and loops with the average value being
SB(ql = 1) = SB(ql = −1) = (7.1 ± 0.3) · 10−3,
where δq1,q2 is the Dirac delta such that δq1,q2 = 1, if
q1 = q2 and 0 otherwise. This sum may be evaluated by
where the given error is the standard deviation over all
samples. This value was obtained for cycle graphs C120 to
9
central loop's winding numbers at balanced fixed points
in dependence of the number of loops added to the graph
k by
Ql,1(k) = {q1 ∈
(cid:21)(cid:12)(cid:12)(cid:64)n1 ∈ N0 s.t. q1 = n1 · k}.
(cid:20)
(cid:99)
−(cid:98) N
4
(cid:99),(cid:98) N
4
In Fig. 5, a stable fixed point that is balanced in the
corresponding cycle graph, but became non-balanced due
to the loops is shown. The figure shows the cycle graph
C28 to which k = 2 loops have been added making the
stable fixed point with q1 = 2 now non-balanced. This
is due to an asymmetry in the phase variables towards
ϑ = 0 making the resulting fixed point non-balanced as
indicated by a non-zero order parameter, Fig. 5 a, (black
square). For this reason, such fixed points need to be
excluded when calculating the balancing ratio.
Using these two results, we are now ready to write
down an estimate for the basin stability in a graph with
loops
b(Gl, k) =
(cid:88)
= AG(k) · (1 − (cid:88)
ϑ∈BGl (N )
q∈Ql,1
SB(q)
SB(ϑ) ≈ (cid:88)
1 − (cid:88)
1√
2πσ2
q1∈Ql,1(k),ql=0
e−q2/(2σ2))
(15)
.
q∈Ql,1
≈ (1 − 1.42 · 10−2)k ·
e−q2/(6.4·10−2·N∗)
√
2π · 3.2 · 10−2 · N∗
Here, N∗ = N − k · m refers once again to the number of
vertices in the central cycle without loops. Analyzing this
expression, one may notice the relationship to the basin
stability of a simple cycle in Eq. 10. This expression dif-
fers from the one for the simple cycle in terms of a factor
accounting for the basin stability in states with non-zero
winding numbers in loops and in terms of several winding
numbers other than the synchronized state being now ex-
cluded from the summations due to the additional sym-
metries induced by the loops.
Importantly, this result
will always yield lower values of basin stability in com-
parison to the corresponding cycle graph. This approx-
imation is shown in Fig. 6 (light purple) for k = 2, 3, 5
along with the corresponding results from Monte Carlo
sampling where the number of loops for the respective
graph is indicated by a purple number.
5. Balancing ratio allows to compare balanced states in
different types of balanced graphs
Finally, the balancing ratio may also be calculated nu-
merically using the Monte Carlo method for different
types of balanced graphs for which there are no ana-
lytical results available so far. To compare the results
for those and the previously discussed topologies among
each other, we plotted the balancing ratio b(G) against
the overall number of vertices in the graph N in Fig. 6.
Different colors and symbols encode different types of
FIG. 5. Winding numbers multiple of the number of loops in-
duce asymmetries in the phase variables at stable fixed points.
(a) Phase variables at the stable fixed point depicted in (b).
The asymmetry towards ϑ = 0 is clearly visible and indicated
by the Kuramoto order parameter in Eq. 3 (black square),
which assumes a non-zero value for this configuration.
(b)
Cycle graph C28 to which k = 2 loops have been added. The
stable fixed point shown here with central winding number
q = 2 is not balanced.
which k ∈ {0, 2, 4, 5, 6, 7, 10, 12, 15, 20, 24, 30} loops have
been added, but was also confirmed for much smaller
cycle graphs C30 with loops added where a similar value
was obtained. Thus, the probability of finding a non-zero
winding number in one of the loops when starting from
a random initial condition for the given graphs reads
p(ql = 1) = 2 · SB(ql = 1) = (1.42 ± 0.06) · 10−2. (13)
This allows us to estimate the average basin stability oc-
cupied by states with a non-zero winding number in de-
pendence of the number of loops k by
AG(k) = (1 − p(ql = 1))k,
(14)
due to the fact that the probability of observing a zero
winding number in a given loop reads 1 − p(ql = 1).
Here, we assumed the probability of observing non-zero
winding numbers in different loops to be independent.
In order to calculate the balancing ratio for this topol-
ogy, we need to derive an expression for the basin stabil-
ity occupied by balanced states. In general, such cycle
graphs with loops allow for a lot of different balanced
states with non-zero winding numbers in some of the
loops according to their symmetry properties, see sec-
tion III B 4 in the appendix for a detailed discussion.
However, in the given setting where loops consist only
of a small number of vertices in comparison to the main
cycle, the basin stability taken by such states is negli-
gible. For this reason, we will focus on balanced states
where all loops have a zero winding number. Compared
to the balancing ratio for the cycle graph, basin stability
for such graphs will then be reduced by the above factor
AG(k).
In addition to that, the symmetry of the added loops
makes a few previously available balanced fixed points
now non-balanced ones. This is the case if the winding
number is a multiple of the number of loops. To formalize
this argument, define the following set characterizing the
+20π2π3π20π2π3π22πab10
FIG. 6. Balancing ratio b(G) plotted for different balanced graphs against total number of vertices in graph N . Color code and
symbols represent different prototypes for the graphs analyzed here as shown in legend, right. For the cycle graph with loops
purple numbers 2, 3 and 5 correspond to the number of loops added. For graphs on the left side of the legend, only the central
cycle grows with increasing N whereas the added loops - if they exist - stay constant in size. For graphs on the right side of the
legend, loops grow accordingly with increasing N (top). Graph at the bottom right of the legend marked by yellow color is the
only asymmetric graph analysed here. Results marked by symbols were obtained using Monte Carlo sampling performed for
respective graph except for analytical results indicated by colored lines. See text for detailed information on how the respective
graph was constructed exactly. Errors take a maximal value of σ(SB(ϑ)) = 2 · 10−3 for each dot and are thus too small to be
visible in the figure.
graphs as indicated by different prototypes shown in the
legend on the right.
The cycle graph appears as an upper bound on the bal-
ancing ratio for a given number of vertices N out of all
balanced graphs. This might be due to the simple struc-
ture of cycle graphs making the balanced states most
easily accessible for random initial configurations and the
fact that there is only one non-balanced fixed point for
this topology.
Furthermore, we note that all of the graphs analyzed
here show a monotonic scaling with the number of ver-
tices in the graph. While most graphs show an increase
in the balancing ratio with the number of vertices in the
graphs, cycles with long loops added and growing in ver-
tices by growing loop vertices as well show a monotonic
decrease with the overall number of vertices, see sym-
bols × and + and light green color in Fig. 6. These
graphs were constructed by adding two loops of equal
lengths l to both sides of a cycle graph symmetrically
with the loops spanning l − 1 of the cycle's vertices
such that the additional cycles now share more than one
edge with the central cycle. For the graphs represented
by symbol ×, loops are of lengths l ∈ {2, 3, 4, 5, 6, 7}
added to the cycle graph C18 resulting in graphs of sizes
N ∈ {22, 24, 26, 28, 30, 32}, respectively. Other copies
of this kind of graph are marked by symbol + and cre-
ated from the cycle graph C46 with higher numbers of
vertices N ∈ {50, 52, 54, 56, 62, 70} and loops of lengths
l ∈ {2, 3, 4, 5, 8, 12} and show similar values of balanc-
ing ratios and a similar scaling. These graphs show a
clear decrease with the length of the loops added which
can be explained by noticing that the length of the loops
corresponds to the number of possible fixed points in the
graphs due to the increase in cycle length. The new fixed
points will occupy some basin stability themselves such
that the balanced fixed points occupy less and less basin
stability with increasing length of loops. This scaling
was confirmed by some preliminary analysis counting the
share of fixed points being balanced. The upper bound
given by the corresponding cycle graphs C46 and C18 is
reached if the length of the loops is reduced to zero.
Furthermore, the graphs represented by symbol (cid:72) and
light green color represent cycle graphs to which more
complicated structures have been added on two symmet-
ric ends resulting in three additional cycles at both sides.
This graph shows the same trend as the simple cycle
graphs to which loops have been added, which are shown
as light purple numbers. This is due to the fact that
these graphs are similar to the latter graphs but different
in terms of the number of cycles in the graph.
20406080100N0.00.20.40.60.81.0b(G)loopsizeconstantloopsizegrowsasymmetricIn terms of the effect additional loops have on balanced
states in cycle graphs, it is easily visible that they result
in a lower balancing ratio compared to the correspond-
ing cycle, see graphs represented by purple numbers 2, 3
and 5 indicating the number of loops. Counterintuitively,
more loops added to the graph might lead to a higher
balancing ratio than fewer loops, although the former
represent a stronger perturbation to the circular topol-
ogy. This is due to the fact that more states with lower
winding number remain balanced if the number of loops
is increased, which take a higher amount of basin stabil-
ity compared to the states with higher winding numbers.
This effect compensates the additional factor AG(k) for
increasing numbers of vertices in the graph leading to
higher balancing ratios for the graphs with more loops.
However, the reduction in the balancing ratio for any
number of loops added symmetrically to the graph is al-
ways smaller than adding a central edge to the graph
which leads to a much stronger decrease in the balancing
ratio, see blue markers (cid:5). All loops added here consist of
m = 3 additional vertices.
For the graph formed by a cycle graph to which a cen-
tral edge and loops have been added, represented by sym-
bol (cid:68) and cyan color, the two small cycles were of size
n ∈ {13, 20, 36} each and k ∈ {2, 3, 4} loops of size m = 3
have been added to each of the cycles resulting in the
points shown at N ∈ {42, 52, 60, 84, 90, 96}, respectively.
In contrast to the results for the single cycle to which
loops have been added, the graphs show similar values of
balancing ratios as for the corresponding graph created
from adding a central edge to the cycle graph.
Finally, the graph represented by yellow color and
symbol (cid:78) is the only asymmetric graph shown here.
This graph is created by adding two edges to the cy-
cle graph C60 asymmetrically, thus resulting in a graph
with cycles of length 31, 21 and 11. The only balanced
fixed points for this graph are characterized by winding
number q = 6 in the corresponding cycle graph.
It
shows a small, non-vanishing balancing ratio of around
b(G) ≈ 3 · 10−3.
IV. SUMMARY AND OUTLOOK
In this work, we showed how additional loops and
edges affect stable, balanced fixed points in the Kuramoto
model on cycle graphs. To do so, we used the roots of
unity as a particular class of balanced states to construct
graphs that support stable balanced states. We exam-
ined the basin stability of stable fixed points in the con-
structed topologies as a measure of stability and used
the results to evaluate the balancing ratio as a new order
parameter for balanced graphs. In order to quantify the
effect of additional connections on balanced states, we
evaluated this parameter both analytically and numeri-
cally for different topologies. Our results show that there
are numerous non-circulant balanced topologies that may
be constructed using simple building blocks. In general,
11
the addition of loops or edges to cycle graphs was shown
to worsen the graph's balancability, with the effect of
adding a central edge to the balanced graph being much
stronger than the one of additional loops.
The balancing ratio introduced here as an order pa-
rameter for balanced graphs provides an easily accessible
tool to classify stability in balanced graphs. We offer
an analytical description of this parameter for a subset
of balanced topologies which might help to yield more
insight into the structure of balanced fixed points in gen-
eral. On the other hand, the balancing ratio may be eas-
ily calculated for any topology using numerical methods.
It also offers the simple interpretation of the probability
of ending up in a balanced state. This could be rele-
vant for real-world networks if balanced states describe a
desired mode of operation.
Using two basic building blocks, we showed how a large
class of balanced graphs may be constructed. In princi-
ple, the building blocks shown here might also be used
to create graphs representing tilings of two dimensional
space but still being balanced. An important task for fu-
ture work would be to find a classification for the building
blocks on more mathematical grounds and to check which
planar balanced graphs exist that may not be produced
using the above schemes. Furthermore, one might move
away from planar graphs and graphs with small degree
being the main focus in this work and look for classifica-
tion schemes for balanced graphs in general.
We found that the standard deviation of basin stabil-
ity in terms of the winding numbers scales linearly with
the number of vertices in the graph for many topologies.
This result yields new insights about the winding number
distribution of stable states in cycle graphs and beyond.
However, it is unclear how the prefactor 3.2 · 10−2 found
to describe the slope of the variance σ2 in terms of the
number of vertices in a cycle graph N may be explained
theoretically. Neither is there a theoretical explanation
for the probability of finding a state with non-zero wind-
ing number in one of the loops p(ql = 1) nor a scaling
of this number with number of loops k, size of loops m or
number of vertices N . One possible way to continue the
study of these graphs would be to quantify this probabil-
ity for different sizes of loops and relate them to the Gaus-
sian scaling observed for connected cycles of the same size
or single cycles. On the other hand, it would be interest-
ing to see if the linear scaling of the Gaussian's standard
deviation with the number of vertices found here to be
valid for two circulant graphs extends to other circulant
graphs - or even other topologies. Using this scaling and
combining different graphs, it could be possible to ana-
lytically calculate the basin stability of single fixed points
for much larger graphs.
Although most of the networks constructed show an
obvious symmetry in correspondence with the roots of
unity, we were not able to relate our findings to results
on cluster synchronization in coupled oscillator models
in relationship to graph symmetry [36, 37] or symme-
tries in the master stability function [38]. It would be an
important goal for future works to be able to relate the
results found here to graph symmetries in the underlying
networks.
The present work is the first one to study planar graphs
under global constraints that manifest through phase bal-
ancing of oscillators on more general grounds. The build-
ing blocks studied here allow to create large networks
of oscillators supporting balanced states, providing the
mathematical framework to understand constrained os-
cillators in nature but also to encode control in robotics
12
and autonomous vehicles. The methods introduced here
further allow to quantify the stability of the balanced
states in such networks - for many networks even analyt-
ically.
ACKNOWLEDGMENTS
This work was funded by the Max Planck Society.
[1] A. T. Winfree, The Geometry of Biological Time, 2nd ed.
(Springer Berlin Heidelberg, 2010).
[2] A. Pikovsky, M. Rosenblum, and J. Kurths, Synchroniza-
tion: A Universal Concept in Nonlinear Sciences (Cam-
bridge University Press, Cambridge, England, 2003).
[24] A. Jadbabaie, N. Motee, and M. Barahona, in Proceed-
ings of the 2004 American Control Conference, Vol. 5
(2004) pp. 4296 -- 4301.
[25] P. Hadley and M. R. Beasley, Appl. Phys. Lett. 50, 621
(1987).
[3] A. Arenas, A. D´ıaz-Guilera, J. Kurths, Y. Moreno, and
[26] S. Nichols and K. Wiesenfeld, Phys. Rev. A 45, 8430
C. Zhou, Phys. Rep. 469, 93 (2008).
(1992).
[4] J. A. Acebr´on, L. L. Bonilla, C. J. P´erez-Vicente, F. Ri-
tort, and R. Spigler, Rev. Mod. Phys. 77, 137 (2005).
[27] T. Y. Lam and K. Leung, J. Algebra 224, 91 (2000).
[28] D. Manik, M. Timme,
and D. Witthaut, Chaos 27,
[5] S. H. Strogatz, Physica D 143, 1 (2000).
[6] Y. Kuramoto, Chemical Oscillations, Waves, and Turbu-
083123 (2017).
[29] F. Dorfler, M. Chertkov, and F. Bullo, Proc. Natl. Acad.
lence (Springer Berlin Heidelberg, 1984).
Sci. USA 110, 2005 (2013).
[7] F. C. Hoppensteadt and E. M. Izhikevich, Weakly Con-
[30] R. Delabays, T. Coletta, and P. Jacquod, J. Math. Phys.
nected Neural Networks (Springer, 2012).
57, 032701 (2016).
[8] E. M. Izhikevich and Y. Kuramoto, in Encyclopedia of
Mathematical Physics (Academic Press, Oxford, 2006)
pp. 448 -- 453.
[9] F. Dorfler and F. Bullo, Automatica 50, 1539 (2014).
[10] R. Sepulchre, D. A. Paley, and N. E. Leonard, "Group
coordination and cooperative control of steered particles
in the plane," in Group Coordination and Cooperative
Control (Springer Berlin Heidelberg, 2006) pp. 217 -- 232.
[11] D. A. Paley, Cooperative control of collective motion for
ocean sampling with autonomous vehicles, Ph.D. thesis,
Princeton University (2007).
[12] D. A. Paley, N. E. Leonard, R. Sepulchre, D. Grunbaum,
and J. K. Parrish, IEEE Contr. Syst. Mag. 27, 89 (2007).
[13] J. Jeanne, N. E. Leonard, and D. Paley, in Proceedings
of the 44th IEEE Conference on Decision and Control
(2005) pp. 3929 -- 3934.
[14] K. Alim, G. Amselem, F. Peaudecerf, M. P. Brenner,
and A. Pringle, Proc. Natl. Acad. Sci. USA 110, 13306
(2013).
[15] L. Scardovi, A. Sarlette, and R. Sepulchre, Systems &
Control Letters 56, 335 (2007).
[16] C. Godsil and G. Royle, Algebraic Graph Theory
(Springer New York, 2001).
[17] R. Diestel, Graph Theory (Springer, 2017).
[18] R. M. Gray, Found. Trends Commun. Inf. Theory 2, 155
(2006).
[19] R. Sepulchre, D. A. Paley, and N. E. Leonard, in in Proc.
Int. Symp. NOLTA (2005).
[20] S. Watanabe and S. H. Strogatz, Physica D 74, 197
(1994).
[21] P. J. Menck, J. Heitzig, N. Marwan, and J. Kurths, Nat.
Phys. 9, 89 (2013).
[22] D. A. Wiley, S. H. Strogatz, and M. Girvan, Chaos 16,
015103 (2006).
[23] H. K. Khalil, Nonlinear Systems, 3rd ed. (Pearson, 2001).
[31] R. Delabays, T. Coletta, and P. Jacquod, J. Math. Phys.
58, 032703 (2017).
[32] L. DeVille and B. Ermentrout, Chaos 26, 094820 (2016).
[33] H. Weyl, Math. Ann. 71, 441 (1912).
[34] R. Delabays, M. Tyloo,
and P. Jacquod, Chaos 27,
103109 (2017).
2πσ)−1(cid:82) q+0.5
the Gaussian SB(q)
q−0.5 e−x2/(2σ2) dx as is done in Ref.
[35] Alternatively, one might calculate the variance by
!=
√
fitting the integral under
[22].
(
However, we found the given approach to yield better
results when comparing it to results from Monte Carlo
sampling.
[36] L. M. Pecora, F. Sorrentino, A. M. Hagerstrom, T. E.
Murphy, and R. Roy, Nat. Commun. 5, 4079 (2014).
[37] M. T. Schaub, N. O'Clery, Y. N. Billeh, J.-C. Delvenne,
and M. Barahona, Chaos 26, 094821
R. Lambiotte,
(2016).
[38] T. Dahms, J. Lehnert, and E. Scholl, Phys. Rev. E 86,
016202 (2012).
Appendix A: Non-symmetric balanced networks
1. Graph automorphisms and symmetries
An important tool to study dynamical systems evolv-
ing on graphs - and in particular balanced states as
possible dynamics - are graph symmetries. Graph au-
tomorphisms in this section being the major tool used
for studying symmetries on graphs. We will introduce
them following Ref. [17, ch. 1]. Consider two graphs
G and G(cid:48) with vertex sets V (G), V (G(cid:48)) and edge sets
13
FIG. 8. A balanced graph with empty automorphism group
may be created by combining the two building blocks of
adding loops to cycle graphs and connecting copies of the
resulting graphs.
(a) A balanced graph with N = 12 ver-
tices and three symmetrically placed loops. (b) A balanced
graph with N = 36 vertices created by joining together three
copies of the graph in (a) resulting in a graph with empty
automorphism group.
graphs.
Appendix B: Set of balanced states for N < 5
In this section, we show that for N < 5, all balanced
states may be constructed making use of the roots of
unity.
Consider the case N = 2. We will show that the
balanced states are given by the second roots of unity,
i.e. that B(2) =(cid:8)ϑ1 = α, ϑ2 = α + π, α ∈ S1(cid:9).
Proof. We have
eiϑ1 + eiϑ2 = 0 ⇒ ei(ϑ2−ϑ1) = −1
⇒ ϑ2 − ϑ1 = π,
This means that in general, we have ϑ1 = α1 ⇒ ϑ2 =
α1 + π, which are the angles building the second roots of
unity.
Along the same lines, the set of balances states for
(cid:8)ϑj = 2πj
3 + α(t), α(t) ∈ S1, j ∈ {1, 2, 3}(cid:9),
N = 3 is given by
B(3)
i.e. the third roots of unity.
=
Proof. Fixing ϑ1 = 0 without loss of generality, we get
eiϑ2 + eiϑ3 = −1
⇒ sin(ϑ2) + sin(ϑ3) = 0 ∧ cos(ϑ2) + cos(ϑ3) = −1
⇒ (ϑ2 = −ϑ3 ∨ ϑ2 = π + ϑ3) ∧ cos(ϑ2) + cos(ϑ3) = −1.
Only the first solution, namely ϑ2 = −ϑ3, is compat-
ible with the second equation as ϑ2 = π + ϑ3 yields
cos(ϑ2) + cos(ϑ3) = cos(π + ϑ3) + cos(ϑ3) = cos(ϑ3) −
cos(ϑ3) = 0 (cid:54)= −1. Thus we arrive at 2 cos(ϑ2) = −1
which implies ϑ2 = ± 2π
3 , which are the
angles corresponding to the third roots of unity.
3 , i.e. ϑ3 = ∓ 2π
Finally, a similar result can be proven for N = 4. We
will show that the balanced states are given by pairs of
FIG. 7. Automorphisms for a cycle-like graph with k = 5
loops symmetrically added to the graph and a total of N =
25 vertices. Numbers on vertices (light blue circles) label
vertices uniquely which are connected through edges (dotted,
black lines). The graph displayed (a) is modified by applying
two different automorphisms, namely the two basis elements
of its automorphism group.
(b) The first basis element is
a reflection along a symmetry axis of the graph (straight,
dotted line, reflection indicated by arrow). (c) The second one
is a rotation (grey, dotted arrow) by a length corresponding
to the distance between individual loops. Together, these
symmetries represent the dihedral symmetries.
E(G) and E(G(cid:48)). A map ϕ : V → V (cid:48) between their ver-
tex sets is a homomorphism between the two graphs if it
preserves the adjacency of vertices (vi, vj) ∈ E(G) ⇒
(ϕ(vi), ϕ(vj)) ∈ E(G(cid:48)).
It is an isomorphism if the
opposite is true as well, i.e. the map is bijective and
its inverse is a homomorphism as well, which implies
(ϕ(vi), ϕ(vj)) ∈ E(G(cid:48)) ⇒ (vi, vj) ∈ E(G). Finally, an
automorphism is an isomorphism from G to itself. The
set of all automorphisms of a graph forms a group. Since
automorphisms are bijective, the combination of two au-
tomorphisms is an automorphism once again and the map
preserving each vertices' position, i.e. the identity map,
is an automorphism itself which shows their group struc-
ture.
The main graphs of interest in this publication are
cycle graphs and its modifications. The basis elements
forming the automorphism group of the cycle graph with
loops added symmetrically are shown in Fig. 7. Panel
(a) shows the cycle graph C20 to which k = 5 loops of
size m = 1 have been added symmetrically. Panel (b)
and (c) show the basis elements of the graph's automor-
phism group, namely an elementary reflection along one
of the graph's symmetry axis (b) and an elementary ro-
tation corresponding to the distance between different
loops (c).
a. Non-symmetric balanced graphs
Besides the types of graphs discussed in section III A,
the building blocks may be used to construct balanced
graphs with empty automorphism group,
graphs
without any graph symmetries. This can easily be
achieved by combining the building blocks in such a way
that the added graphs do no longer have trivial symme-
try properties. A simple example obtained by connecting
three cycle graphs with loops in an asymmetric way is
shown in Fig. 8.
In general, this shows that automor-
phism groups alone cannot be used to classify balanced
i.e.
0π2π3π22π⇒ab(cid:110)
second roots of unity, i.e. that
(cid:111)
∧ indices non-equal
.
B(4) =
ϑj = ϑk + π, ϑl = ϑm + π; j, k, l, m ∈ {1, 2, 3, 4}
Proof. We have the following requirements on the angular
variables
eiϑ1 + eiϑ3 = −(eiϑ2 + eiϑ4 )
(P1)
e−iϑ1 + e−iϑ3 = −(e−iϑ2 + e−iϑ4) .
(P2)
Now assume without loss of generality that eiϑ1 +eiϑ3 (cid:54)= 0
and eiϑ1 + eiϑ4 (cid:54)= 0. If one of the two equations held, we
would have ϑ3 − ϑ1 = π or ϑ4 − ϑ1 = π, respectively,
using the proof for the second roots of unity and would
thus be done. Now one can make use of the equality
e−iϑ1 + e−iϑ3 = (eiϑ1 + eiϑ3)/(eiϑ1 · eiϑ3) to obtain the
following result from equation (P2)
eiϑ1 + eiϑ3
eiϑ1 · eiϑ3
=
eiϑ2 + eiϑ4
eiϑ2 · eiϑ4
.
Together with equation (P1), this implies that
ei(ϑ2+ϑ4) = ei(ϑ1+ϑ3)
must hold since eiϑ1 + eiϑ3 (cid:54)= 0. Along the same lines,
one may show that
ei(ϑ1+ϑ4) = ei(ϑ2+ϑ3)
is true. Multiplying these two equations with each other
yields the identities e2iϑ2 = e2iϑ1 and e2iϑ3 = e2iϑ4. This
implies the two equalities ϑ2 = ϑ1∨ϑ2−ϑ1 = π and ϑ4 =
ϑ3∨ϑ4−ϑ3 = π. Only the latter possibilities ϑ2−ϑ1 = π
and ϑ4 − ϑ3 = π are compatible with equations (P1) and
(P2) which completes the proof.
Appendix C: Summation over roots of unity
The following theorem states that N th roots of unity
sum up to zero.
Theorem C.1 (Summation of roots of unity). Let N ∈
N be a natural number and consider the vectors of N th
roots of unity rN,k = eikθN , where θN = (θ1, ..., θN ) and
θj = arg(ρj,N ) = 2πj/N ∈ S1 are angular variables and
k ∈ N<N is a natural number. Then for any natural
number k ∈ N the sum of the roots of unity vanishes
N(cid:88)
j=1
1T rN,k =
eikj2π/N = 0.
14
FIG. 9. Sketch of a cycle graph with k = 6 loops. The
winding numbers in the loops for balanced fixed points are
not limited to be all-zero (top left) or all unity (bottom left,
unity value indicated by colour), but also patterns where the
loop winding numbers are prime divisors of k may result in
balanced fixed points (centre and right plots).
(cid:80)N
j=1 eik(j+1)2π/N . Now one can solve for 1T rN,k, since
k < N and thus 1 − eik2π/N (cid:54)= 0 obtaining
1T rN,k =
(cid:80)N
j=1 eikj2π/N −(cid:80)N
= eik2π/N 1 − eik2π
(1 − eik2π/N )
1 − eik2π/N
= 0,
j=1 eik(j+1)2π/N
where eik2π = 1, ∀k ∈ Z was used in the last step.
Appendix D: Balanced states in cycle graphs with
loops
In this section, we want to discuss the balanced states
in cycle graphs with loops in more detail. Here, we will
focus on the case where loops consist of m = 3 vertices
added to the graph for all loops such that the winding
number of the loops characterizing stable fixed points
reads ql,i ∈ {−1, 0, 1}. To characterize these winding
numbers, we will make use of results on the geometry
of balanced states discussed in section II C. As an exam-
ple, consider a cycle graph with k = 6 loops.
In this
case, states with ql ∈ {(1, 0, 0, 1, 0, 0)T , (1, 0, 1, 0, 1, 0)T}
representing the loop number's prime factors can lead to
balanced states as indicated in Fig. 9. To characterize
these states on more mathematical grounds and express
their relationship to the central loop's winding number
q1, lets define the set of factors of the number of loops k
whose factors are not a divisor of some integer q by
RQl (k, q) = {p1 ∈ N<k(cid:12)(cid:12)∃p2 ∈ N<k : p1 · p2 = k ∧ p2 (cid:45) q},
tion yields (1 − eik2π/N )1T rN,k = (cid:80)N
Proof. Multiplying both sides of the equation by eik2π/N
and subtracting the result from the original equa-
j=1 eikj2π/N −
where the expression p1 (cid:45) q expresses the fact that p1 is
not a divisor of q. Using this idea, one may write down
the set of all loops' winding numbers leading to balanced
15
states. These states occur if the factor between the num-
ber of loops k and the overall number of vertices in the
new graph N∗ = N + k · m is a multiple of the number
of edges added in each loop N∗/k = (m + 1) · p, p ∈ N.
They are characterized by all equal, non-zero winding
numbers in all loops ql,j (cid:54)= 0, ∀j, and a non-zero wind-
ing number q1 in the central cycle with the same sign
sign(q1) = sign(ql,j) which reads q1 = sign·(ql,j)(p−1)·k,
see for example Fig. 3 b.
states
l,2 (q, k) = {ql ∈ D>0
Q>0
l,2 (ql)(cid:12)(cid:12)
(cid:124) (cid:123)(cid:122) (cid:125)
ql ∈ span(σp((1, 0, 0, ..
(cid:124) (cid:123)(cid:122) (cid:125)
p ∈ {1, ..., p1 − 1}), p1 ∈ RQl (k, q)},
, ..., 1, 0, 0, ..
p1
p1
)T ),
where σp denotes again the cyclic shift applied p times
and the >0 superscript indicates that this set is restricted
to positive winding numbers. Note that this set only con-
tains symmetric states if the integer p1 is an integer di-
visor of the number of loops k such that p1
wise RQl is the empty set. The cyclic shift represents the
graph's rotational symmetry. In general, the sign of the
loops' possible winding numbers needs to coincide with
the sign of the central cycle's winding number. Thus,
one may classify the balanced fixed points with positive
winding numbers by
(cid:12)(cid:12)k ∈ N, other-
(k) = {q ∈ Dl(q)(cid:12)(cid:12)q1 ∈ Q>0
l
Q>0
Here, Qc = {q ∈ Dc(q)(cid:12)(cid:12)q1 = q2, q1 (cid:54)= 0} is the set defin-
c (k), ql ∈ Q>0
l,2 (q1)}.
ing the possible winding numbers in the inner cycle as
defined in section III B. The set of negative winding num-
bers may be defined correspondingly and represents bal-
anced states as well. Note that we now take the set of
loop winding numbers in dependence of the winding num-
ber in the inner cycle q1 Q>0
l,2 (q1) such that the main cy-
cle's winding number induces the possible factors for the
loops' winding numbers and thus their symmetry prop-
erties. The cases where p2q1 in the set of factors RQl in-
duces again a global symmetry similar to the one shown
in Fig. 9 b and the resulting states are not balanced. Note
that this definition only characterizes balanced states for
m < 5 and needs to be redefined otherwise.
In some cases, the combination of positive main wind-
ing number and negative loop winding numbers q1 ∈
c (k), ql ∈ Q<0
Q>0
l,2 (q1) and vice versa leads to balanced
states as well, that do, however, not always exist. In most
cases, these states do exist for q1 = 1. Nonetheless, these
fixed points might possess phase differences larger than
π/2, in particular for small graphs. Take for example the
cycle graph C40 to which k = 4 loops have been added.
Then the state characterized by q = (1,−1, 0,−1, 0)T is
stable and balanced, but has two phase differences larger
than the threshold ϕ ≈ π/2 + 0.141 > π/2. For this rea-
son, these states may not be studied using the algorithm
for finding stable, balanced states, but will be accounted
for by the Monte Carlo method if they take a significant
amount of the system's basin stability. However, most
of the states with non-zero loop numbers occupy only a
small share of basin stability in practice and may thus be
neglected when calculating the graph's balancing ratio.
Even though almost complete, this set still doesn't
fully account for the balanced states. The stable fixed
points that appear through zero-energy edges in higher
order winding numbers of the corresponding cycle graph
need to be included when summing over all balanced
|
1510.07259 | 1 | 1510 | 2015-10-25T15:01:27 | Long-range Acoustic Interactions in Insect Swarms: An Adaptive Gravity Model | [
"physics.bio-ph",
"cond-mat.stat-mech",
"nlin.AO"
] | The collective motion of groups of animals emerges from the net effect of the interactions between individual members of the group. In many cases, such as birds, fish, or ungulates, these interactions are mediated by sensory stimuli that predominantly arise from nearby neighbors. But not all stimuli in animal groups are short range. Here, we consider mating swarms of midges, which interact primarily via long-range acoustic stimuli. We exploit the similarity in form between the decay of acoustic and gravitational sources to build a model for swarm behavior. By accounting for the adaptive nature of the midges' acoustic sensing, we show that our "adaptive gravity" model makes mean-field predictions that agree well with experimental observations of laboratory swarms. Our results highlight the role of sensory mechanisms and interaction range in collective animal behavior. The adaptive interactions that we present here open a new class of equations of motion, which may appear in other biological contexts. | physics.bio-ph | physics | Long-range Acoustic Interactions in Insect Swarms: An Adaptive Gravity Model
Dan Gorbonos,1 Reuven Ianconescu,1, 2 James G. Puckett,3 Rui Ni,4 Nicholas T. Ouellette,5 and Nir S. Gova1
1Department of Chemical Physics, The Weizmann Institute of Science, P.O. Box 26, Rehovot, Israel 76100
2Shenkar College of Engineering and Design, Ramat-Gan, Israel
3Department of Physics, Gettysburg College, Gettsyburg, Pennsylvania 17325, USA
4Department of Mechanical and Nuclear Engineering,
The Pennsylvania State University, University Park, Pennsylvania 16802, USA
5Department of Civil and Environmental Engineering,
Stanford University, Stanford, California 94305, USA
The collective motion of groups of animals emerges from the net effect of the interactions between
individual members of the group. In many cases, such as birds, fish, or ungulates, these interactions
are mediated by sensory stimuli that predominantly arise from nearby neighbors. But not all stimuli
in animal groups are short range. Here, we consider mating swarms of midges, which interact
primarily via long-range acoustic stimuli. We exploit the similarity in form between the decay of
acoustic and gravitational sources to build a model for swarm behavior. By accounting for the
adaptive nature of the midges' acoustic sensing, we show that our "adaptive gravity" model makes
mean-field predictions that agree well with experimental observations of laboratory swarms. Our
results highlight the role of sensory mechanisms and interaction range in collective animal behavior.
The adaptive interactions that we present here open a new class of equations of motion, which may
appear in other biological contexts.
5
1
0
2
t
c
O
5
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
9
5
2
7
0
.
0
1
5
1
:
v
i
X
r
a
a Correspondence and requests for materials should be addressed to N.S.G. (email: [email protected]).
I.
INTRODUCTION
2
Collective behavior of groups of social animals is widespread in nature [1], and occurs on size scales ranging from
single-celled organisms [2, 3] to insects [4 -- 7] to larger animals such as birds [8, 9] or fish [10 -- 12]. Animals are thought to
aggregate and move cooperatively for many reasons; collective behavior may, for example, reduce the risk of predation
for an individual in a group [1, 13], promote efficient mating and decrease inbreeding in dispersed populations [14],
modulate the energetic cost of migration [15], or enable enhanced sensing [12]. Because of both its ubiquity and the
potentially advantageous properties it conveys for groups, collective behavior has engaged a broad cross-section of
scientists, ranging from physicists and applied mathematicians who hope to tease out the general principles that drive
the emergence of collective states in non-equilibrium systems to engineers who hope to develop bio-inspired control
strategies for distributed multi-agent systems.
Collective behavior therefore has a long modeling history [1, 16, 17]. Models are useful both as a check on our
fundamental understanding of the kinds of low-level interactions that lead to the emergent group properties and as a
stepping stone to the design of engineered systems that exploit them. Many models treat the group as a collection
of self-propelled agents that are in some way coupled [18, 19]. Building such an agent-based model therefore requires
several fundamental choices [20]. We must specify the base-case, non-interacting behavior of each individual; we must
choose a functional form for the interactions that couple the individuals; we must decide whether these rules are
uniform throughout the population and in time; and we must decide which individuals interact. Each of these choices
can be difficult to make with certainty, and yet has significant ramifications for model performance and fidelity.
Here, we focus on the last of these modeling assumptions: the choice of which individuals interact. From passive
observations alone, it is difficult to discern the correct interactions in a group of animals [5], since it requires the solution
of a challenging inverse problem. Thus, it is common to replace the difficult-to-measure interaction network with the
more straightforward proximity network [21]; that is, one assumes that the local neighborhood of an individual
dominates that individual's behavior. This local neighborhood may be defined by, for example, either metric or
topological distance [22, 23], but it is still the local environment that is assumed to matter. This assumption is
reasonable, and appears to be valid [10, 11, 22, 24, 25], for dense groups of animals such as flocks of birds or schools of
fish that interact primarily through vision and that move in a coordinated, directed fashion. But it need not always be
true; crowds of humans moving toward goals, for example, can show emergent collective behavior while "interacting"
with other individuals with whom they are likely to collide in the future rather than those who are closest to them
[26].
We consider here a canonical example of collective animal behavior -- mating swarms of flying insects -- where local
interactions do not clearly play a major role and yet where the animals display group-level cohesion [5]. Recently
this system has also generated interest for possible indications of critical behavior [7, 27]. Previous descriptions of
insect swarms have accounted for the tight binding of individuals to the group either by introducing a confining
potential [6, 7] or by invoking external environmental cues [14]. Here, we instead develop a swarm model inspired by
the dominant sensory mechanism of the insects, and show that group cohesion can emerge naturally instead of being
externally imposed. Swarming species of Chironomid midges, such as those we consider here, are known to be very
sensitive to acoustic signals [28], and are thought to be attracted to swarms by the sound they produce.
We thus make the ansatz that midges accelerate toward the sound produced by others; in essence, we hypothesize
that in addition to their known pairwise function, acoustic interactions are the basis for coordinating the large-scale
collective behavior of the swarm. The exact structure of the acoustic field produced by a freely flying midge is
not known, and to our knowledge there are no direct measurements of its detailed spatial structure. However, in
other flying insects (flies, for example) the acoustic field produced has been found to have both monopole and dipole
components [29]. Since the monopole field decays more slowly compared to the dipole (and any higher multipole)
component, in the model we present we have concentrated on its contribution. The monopole sound intensity falls off
according to an inverse-square law, and so this hypothesis results in a effective gravity-like force that promotes group
cohesion while still allowing for complex individual motion.
We additionally account for the possibility that the midges' sensory perception may adapt to the overall sound level
they experience. Our model is thus not derived from an underlying kinetic theory, as is commonly done in models
of collective behavior [18], but rather is explicitly long-range and (because of the adaptivity) many-body; and, as
the model bears some similarity to self-gravitating systems, we can exploit our intuition for gravity in analyzing our
model.
To benchmark this model, we compare some of its predictions with laboratory measurements of several hundred
swarms of the non-biting midge Chironomus riparius, and find surprisingly good agreement. Since the acoustic
interactions in our model are long-range and adaptivity renders them inherently many-body, our results suggest that
the midges process more than just local information, as has also recently been proposed for bird flocks [30]. We use in
this work concepts and techniques from classical N -body self-gravitating systems, to explain the collective swarming
behavior of insects (midges). Although models of collective behavior abound in the literature, this particular approach
3
has never been applied to the swarming problem. What we find especially appealing about it is both that a gravity-like
model can produce very complex behavior from simple interactions, as is expected to occur in collective behavior in
biology. At the same time, we introduce a concept to the gravitational physics community that is taken from biology,
namely the adaptivity of the sensory mechanism, ending up with a new form of gravitational interaction that we term
"adaptive gravity".
II. EXPERIMENTAL SETUP
Before discussing our model and its comparison with our empirical data, we describe our laboratory experiments
with midge swarms. The details of these measurements have been discussed elsewhere [4, 5, 31], and so we only give
a brief overview here.
Our self-sustaining colony of C. riparius midges was originally established from egg sacs purchased from Environ-
mental Consulting and Testing, Inc. The colony is maintained in a transparent cubical enclosure, 91 cm on a side, at
a constant 22◦C. The midges are exposed to overhead light on a circadian cycle, with 16 hours of light and 8 hours
of darkness per day. When the overhead light turns on and off (corresponding to "dawn" and "dusk"), adult males
spontaneously form swarms. To promote swarm nucleation and to position the swarm in the enclosure, we use a
black felt "swarm marker" measuring 30 × 30 cm2 placed just above the development tanks. Further details of our
husbandry procedures are given in refs. [4] and [5].
To quantify the motion of the midges, we record movies of swarming events with three hardware-synchronized Point
Grey Flea3 cameras at a rate of 100 frames per second. We have shown previously that this data rate is sufficient to
resolve even the acceleration of the midges [4]. The cameras are arranged in a horizontal plane about 1 m from the
center of the swarm with angular separations of approximately 30◦ and 70◦. Prior to recording data, the camera system
is calibrated using Tsai's model [32]; subsequently, the two-dimensional coordinates of each midge on each camera
(found by simple image segmentation and intensity-weighted averaging) can be combined to find the midge positions
in three-dimensional space. The sequences of time-resolved positions are then linked into trajectories using a fully
automated multi-frame predictive tracking algorithm [33]. For various reasons, trajectories may sometimes be broken;
thus, in a post-processing step we link trajectory fragments using Xu's method of re-tracking in a six-dimensional
position-velocity space [34]. After trajectory construction, we compute accurate time derivatives by convolving the
tracks with a smoothing and differentiating kernel [4].
For the results shown here, we analyzed data from 128 swarming events. Although the number of individuals was
not uniform from swarm to swarm, we have shown previously that the swarms reach a statistical "asymptotic" regime
at surprisingly small population sizes [31]; here, the mean number of individuals per swarm was about 10. Finally,
for reference below, we note that the body size of a male C. riparius midge is about 7 mm in length. Typical flight
speeds of the midges are roughly 0.5 m/s, and peak instantaneous accelerations are on the order of 5 m/s2. The sound
produced by a male's beating wings is broadband, but has a fundamental frequency of about 575 Hz, as measured
in our experiments [35]. It is difficult to measure the sound amplitude produced by a freely flying midge precisely;
within a few body lengths of the midge, we have measured it to be roughly 55 dB.
In previous work, we have analyzed data from these swarms to characterize their dynamics [4, 5, 31, 36]. Without
going into detail here, we briefly summarize our primary findings. Although our midge swarms remain confined to a
compact region of space with a statistically sharp boundary (in a way that appears to be self-organized [31]), teasing
out pairwise interactions within the swarm is very challenging [5, 36] and the swarms show no net internal order [4].
At a mean-field level, the statistics of the swarms share some features of an ideal gas in a harmonic trap [4, 5]. Thus,
with our model, we attempt to capture these primary effects: strong binding of individuals to the swarm as a whole
but no strong signature of pairwise interaction at the mean-field level, overall disorder inside the swarm, and complex
individual trajectories.
III. MODEL
In the absence of viscous damping, the intensity of the monopole component of the sound produced by an isotropic
point emitter decays purely geometrically: since the acoustic energy in the spherical wavefront is fixed and the surface
area of these wavefronts grows as r2, where r is the distance from the emitter, the sound intensity must fall off as
1/r2. The typical fundamental frequency produced by the wingbeats of a male midge in our experiments is roughly
575 Hz [35], corresponding to a wavelength of about 60 cm. Given the typical size of a swarm (diameters of roughly
20 cm in our experiments) and nearest-neighbor distance (about 4 − 7 cm [4]), midges in a swarm are therefore in the
near-field regime of the acoustic field of their neighbors. In refs. [37, 38], the decay of the pressure field due to the
wing-beat of a fly was measured, and it was found that both in the near-field and far-field the decay follows a 1/r
4
behavior. This results in a decay of the intensity as 1/r2, which is the form we use in our model. We note that due to
interference effects between the incoherent sources, the acoustic field decays roughly as a Gaussian for length scales
larger than the wavelength, i.e. & 60 cm; this size is, however, much larger than the typical swarms we have studied.
Next, we make the hypothesis that an individual midge accelerates towards a neighbor via an effective "force"
that is proportional to the sound intensity. Given the estimates above, this means that the force between a pair of
midges i and j separated by a distance rij = ~ri − ~rj will scale as 1/r2
ij, just as the gravitational attraction between
a pair of point masses would. At present, we must treat this choice purely as an ansatz, as the details of the form of
any pairwise interactions between midges is very difficult to access experimentally [5, 36]. However, the assumption
the midge response to acoustic signals is an acceleration towards the sound source is the simplest choice one can
make. Choosing the response to be at the velocity level would be somewhat unnatural, since velocity cannot be
directly controlled by the insects: changes to the velocity must come from forces applied by the insect, and therefore
accelerations. And strong (albeit indirect) experimental support for this assumption comes from the observation of a
net linear restoring force acting towards the swarm center (Fig. 1b). The only form of binary interactions that gives
this linear restoring force towards the swarm center is an inverse-square force relation.
For many animals, the perception of sound is not fixed, but rather adapts to the total sound intensity so that
acoustic sensitivity drops when there is strong background noise. This is a common feature of biological sensory
organs, preventing their damage and saturation. We thus make a second ansatz: that in general the midges' acoustic
perception adapts to the overall sound level, and that specifically it follows the fold-change detection mechanism [39],
which is ubiquitous in nature. In that case, the effective force on midge i due to midge j is given by
~F i
ef f = CXj
rij
1
~ri − ~rj2(cid:18)
R−2
R−2
ad
ad +Pk ~ri − ~rk−2(cid:19) ,
(1)
where rij is the unit vector pointing from midge i to midge j, C is a constant with dimensions of mass·length3/time2,
and Rad is the length scale over which adaptivity occurs. In other words, when a single midge is closer than Rad the
sound it emits is strong enough that the receiving midge needs to adapt its sensitivity to reduce the perceived signal.
Beyond this distance there is no need for such adaptivity for the sound of a single midge.
Equation (1) constitutes the core of our "adaptive gravity" model (AGM) for the acoustic interactions of the midges.
We note that this model assumes that the midges can sense both the intensity and the direction of the sound produced
by others; it is thought, however, that the specialized Johnston's organs of male swarming insects are indeed able to
do so [40]. We also note that we are making the simplifying assumption that each midge is identical; although this
assumption is certainly not fully accurate, it should allow us to make reasonable mean-field predictions.
With the assumption of adaptivity, the force felt by each midge is inherently many-body and cannot be written as
a sum of two-body interactions (Eqs. (3) and (4)) due to the sum over all the midges that appears in the denominator
of the adaptivity factor (Eq. (1)). Thus, in this formulation, every midge feels a force that contains global, long-range
information about the swarm, but this force cannot be parsed to distinguish the effects of any single neighbor. Thus,
in this AGM the force that binds individual midges to the swarm is truly an emergent, group-level property that
arises naturally from within the swarm without any appeal to external effects.
To build intuition for the behavior of this model, let us consider two limits. For rij ≫ √N Rad (that is, when the
distance between a pair of midges far exceeds the range of adaptivity, and N is the number of midges in the swarm),
the effective force reduces to a purely gravitational interaction and becomes
~F i
ef f,g → CXj
rij
1
~ri − ~rj2 .
(2)
In the opposite limit, when rij < √N Rad, the adaptive nature of the acoustic sensitivity becomes dominant. In that
case, the adaptivity simply reduces to a rescaling of the sound perceived by each midge by the total buzzing noise
amplitude, and the effective force becomes
~F i
ef f,a →
ad ·Pj rij
C
R2
Ntot
1
~ri−~rj2
,
(3)
where the total buzzing noise amplitude at ~ri is Ntot =Pj(~ri − ~rj)−2.
In pure gravity, the potential is additive, and the principle of superposition applies. Due to adaptivity, however,
this property is lost in our model. That is, the effective potential felt by a midge due to many other midges is not the
sum of two-body interactions using Eq. (A3). This can be seen by considering many interacting midges. The effective
force felt by midge i due to the others (indices j) is given in Eq. (1) which is not equal to the sum over two-body
5
Figure 1. Mean-field effective forces. a. Calculated mean effective force acting on a midge within a spherical swarm as a
function of the radial position r. The solid line shows the force due to adaptive gravity (Eq. (1)) for a swarm of uniform
density and radius Rs, compared to the case of pure gravity (dashed line). When the density has a Gaussian profile (with
spatial variance Rs/2, Eqs. (8)-(10)), the adaptive-gravity interactions give rise to the force shown with the dotted line. b.
Measured mean acceleration as a function of position for laboratory swarms with different numbers of midges (shown with
different colors), showing the roughly linear behavior near the swarm center.
forces (see Eq. (A1) in the Appendix), which would be
~F i
ef f,2 = CXj
rij
1
R−2
ad
~ri − ~rj2
R−2
ad + ~ri − ~rj−2
.
(4)
Unlike in pure gravity, the forces on midge i due to others are not additive and the superposition principle does not
apply, since the total buzzing noise term does not depend on direction. As a consequence, there are no conservation
laws in such a system (except mass conservation), and the sum of forces felt by all the midges within an isolated swarm
need not vanish, as it must in regular gravity. Thus, in this model the center of mass of the swarm can experience
accelerations; so, even though the AGM naturally leads to swarm cohesion, one may need to posit external effects to
prevent drift of the swarm as a whole.
IV. RESULTS
Effective spring constants: Spherical swarms.
Gauss's law for gravity states that the gravitational flux through a closed surface is proportional to the enclosed
mass [41]. In our analogy, each midge has an effective unit "mass," and therefore Gauss's law for the force in the
"pure gravity" regime (Eq. (2)) is
Z∂V
~Fef f · d ~A = −4 π C ZV
ρ(~r)d3r,
(5)
where V is a three-dimensional volume, ∂V is its boundary, d ~A is a surface element, and ρ(~r) is the density of midges.
We begin with a spherical swarm of uniform density ρ and radius Rs = hri. The characteristic value of the total
S, where N is the number of midges. Thus, when RS ≫ √N Rad
buzzing noise intensity at the origin is Ntot ∼ N / R2
we are in the pure gravity regime. In this case, from the analog of Gauss's law (Eq. (5)), the force is restoring and
linear with respect to the distance ~r of a midge from the center of the swarm (Fig. 1a), and is given by
~Fef f = −
4 π C ρ
3
~r.
(6)
Since this force is harmonic (that is, restoring and linear in ~r), we can characterize its strength with an effective
"spring constant" K = 4 π C ρ/3. We stress that this behavior is unique for rij /r2
ij interactions, assuming that the
motion arises only from interactions between the midges. Previously, we found that the average acceleration of midges
6
Figure 2. Effective spring constants in the horizontal directions (Kx and Ky) and the vertical direction (Kz) as a function of
the average swarm radius Rs, plotted on logarithmic axes. The black circles denote the raw data for each swarm, the red circles
show binned averages, and the dashed lines denote the R−1
scaling predicted by the AGM (Eq. (7)). We also plot the R−2
scaling predicted for cylindrical swarms (regular gravity, Eq. (A7)), which seems to fit the behavior of larger swarms. Note
that in the vertical z-direction the data is more scattered, which may be due to Earth's gravity.
s
s
in laboratory swarms also has this harmonic form [4], providing support for our model. However, in these laboratory
swarms, the spring constants were found to depend on the swarm size Rs (Fig. 1(b)), unlike in pure gravity.
For swarms with large numbers of individuals, however, our model enters its adaptive regime, where RS < √N Rad.
In this regime, the net force is still linear and restoring; but due to the adaptive terms in Eq. (1), the spring constant
ad Rs)−1. This result
K will depend on the swarm size Rs. To leading order in r/Rs, the model predicts that K ∝ (R2
is derived in the Appendix (Eqs. (A1)-(A3)), and also follows from Eq. (3) using simple dimensional analysis, since
Xj
R2
ad
r2
ij
−1
→ R2
ad Z Rs
0
ij!−1
d3r
r2
∼(cid:0)R2
ad Rs(cid:1)−1
.
(7)
When we examine the experimental data for Kx and Ky (where x and y are in the plane and z is vertical), we find
good agreement with the model prediction that K ∝ R−1
(Fig. 2a,b). This agreement is a consequence of the roughly
constant density in swarms of different sizes (except for small swarms of fewer than ∼ 10 midges [31]), and gives a
lower bound on Rad & 45mm since the adaptive regime applies to the smallest swarms. For Kz we find a decrease
that is faster than predicted (Fig. 2c), as discussed further below.
s
We note that away from the swarm center the adaptive-gravity interaction gives rise to a restoring force that
deviates from the form of pure gravity even for a uniform density swarm (Fig. 1a). Thus, we also calculated the force
for a swarm with a Gaussian density profile, as was observed in experiments [4](Fig. 1a), and find that it is roughly
linear over the entire swarm size, but saturates at large radii.
The calculation of the adaptive force near the swarm center, for a Gaussian density profile, is as follows. We take
a Gaussian density profile with width σs, so that the density in cylindrical coordinates is given by
The gravitational force at z0 is then
ρ(r, z) =
exp (− r2+z2
2 σ2
s
2 σ3
(2 π)
S
3
)
.
Fef f,g(z0) = 2 πZ ∞
−∞
dz′Z ∞
0
r′dr′ρ(r′, z′)
z′ − z0
[r′2 + (z′ − z0)2]
,
3
2
and the total buzzing noise is
Ntot(z0) = 2 πZ ∞
−∞
dz′Z ∞
0
ρ(r′, z′)
r′dr′
r′2 + (z′ − z0)2 .
(8)
(9)
(10)
The integrals were solved numerically using M athematica 9.0, and were used to plot the dotted line in Fig. 1a.
To conclude this part, we find that the accelerations of midges near the swarm center follow the linear relation
expected from gravity-like interactions. Furthermore, the effective spring constant decreases with swarm size, exactly
as predicted by adaptivity. The effects of non-spherical swarms are dealt with next.
7
Figure 3. (a) Observed average inertia eigenvalue ratio I1/I3 as a function of the swarm size Rs (blue circles: raw data for
each swarm, red circles: binned average). For smaller swarms, the ratio is ∼ 1 which is the spherical limit, while for larger
swarms this ratio is larger than 1 due to the elongation of the swarm in the z-direction (I3 > I1, Fig. 4a). (b) Calculated spring
constant ratio K1/K3 for a prolate, axisymmetric swarm, using Eq. (21), as a function of the aspect ratio p ≡ a/c (solid line).
For small p the behavior is linear, as given in Eq. (22).
A. Effective spring constants: Ellipsoidal swarms.
In the measured swarms, the effective spring constant in the vertical (z) direction is consistently smaller than
those in the horizontal (x, y) directions [4]; additionally, it is also observed to decrease faster with swarm size than
is predicted by our AGM for spherical swarms (Fig. 2c). For real swarms, the z direction differs from the x and y
directions in several ways. First, along this direction midges are affected by the earth's gravitational pull. Additionally,
swarms tend to form over visual features on the ground [31], which breaks the isotropic symmetry. Empirically, all
these differences tend to cause larger swarms to elongate along the z axis [4, 31] (Fig. 3a). As we calculate below
(and show in Fig. 3b), for swarms that are elongated along the z-axis, our model predicts that the effective spring
constant in the x, y-plane (K1,2) is larger than in the z-direction (K3). We therefore attribute the observed smaller
spring constant in the z-direction for larger swarms (Fig. 2) to the elongation of the swarms along the vertical axis.
Furthermore, we can calculate the scaling of K3 with swarm size in the limit of a highly elongated (cylindrical) swarm,
such that it has a fixed radius R in the xy-plane and a variable length L ≫ R along the z-axis. In this limit we find
the scaling K3 ∝ R−2
(Eq. (A7)), as denoted in Fig. 2c. Note that in this limit, we are beyond the perfect adaptivity
regime, and the scaling result for K3 is identical to that of pure gravity.
s
In Fig. 4a we show the shape of a typical elongated swarm. We treat the swarm shape using an ellipsoidal
approximation to refine the analysis of the effective spring constants along the different directions. We assume that
the swarm is an ellipsoid with semi-axes a, b, and c, where c < b < a (see Fig. 4a), along the x, y, z-axes. The effective
spring constants are then given by (see Appendix, Eqs. (A1)-(A31))
0
K1 = π a b c ρZ ∞
K2 = π a b c ρZ ∞
K3 = π a b c ρZ ∞
0
0
dv
dv
(c2 + v)pβ(v)
(b2 + v)pβ(v)
(a2 + v)pβ(v)
dv
,
,
,
(11)
(12)
(13)
where β(v) ≡ (a2 + v)(b2 + v)(c2 + v) and K1 > K2 > K3 since c < b < a. Eqs. (11) -- (13) relate the effective forces in
the swarm to its overall shape. The measured values of the spring constants and shapes of the swarms are summarized
in Table 1. We characterize the shapes by the ratios of the moments of inertia-tensor eigenvalues η1 = I1/I2 and
η2 = I1/I3. As part of our ellipsoid approximation, we assume that the inertia eigenvectors are oriented along the
principal axes of the ellipsoid and that each inertia tensor eigenvalue corresponds to one of the axes. Let us assume
8
Figure 4. Swarm shape. a. Example of the shape of a laboratory swarm of N = 7 midges, which is elongated along the vertical
(z) direction. Also shown is the ellipsoid that approximates the swarm shape, which by construction has the same ratios of
inertia eigenvalues. b. The ratio of the effective spring constants Kx,y/Kz as a function of the swarm radius Rs, using the data
in Table 1. Measured values are shown in blue, and those calculated from the model are shown in purple.
that I1 > I2 > I3, without loss of generality. Then
and
I1 =
I2 =
I3 =
4 π
15
4 π
15
4 π
15
ρ a b c(cid:0)a2 + b2(cid:1) ,
ρ a b c(cid:0)a2 + c2(cid:1) ,
ρ a b c(cid:0)b2 + c2(cid:1) ,
η1 ≡
η2 ≡
I1
I2
I1
I3
=
=
a2 + b2
a2 + c2 ,
a2 + b2
b2 + c2 .
From Eqs. (17, 18), one can express the parameters of the ellipsoid as a function of η1 and η2:
b2
c2 =
a2
c2 =
η1 + (η1 − 1) η2
η1 + η2 − η1 η2
η1 − (η1 + 1) η2
η1 (η2 − 1) − η2
,
,
(14)
(15)
(16)
(17)
(18)
(19)
(20)
and then the ratios K1/K2 and K1/K3 can be obtained from Eqs. (11)-(13). Note that in this analysis (Table 1) the
direction of each Ki can be different, as they are defined by their relative strength so that K1 > K2 > K3. In all
cases, however, the smallest effective spring constant is in the z direction (K3 = Kz), since we always observe that
swarms are stretched in the vertical direction (Fig. 3).
The ellipsoid parameters can be expressed in terms of the ratios of the inertia-tensor eigenvalues (Eqs. (14)-(20)),
thereby relating the effective spring constant ratios K1/K2 and K1/K3 to the measured shape. We plot these ratios
for both measured swarms and for the model in Fig. 4b. For smaller swarms, we find good agreement between the
theoretical and measured values, where the discrepancies are primarily due to misalignment of the inertia-tensor
eigenvectors with the principal axes of the ellipsoid. For the largest swarms, however, there is a significant deviation
between the two. In those cases, the ellipsoidal approximation may not be valid, as we sometimes observed a tendency
for these large swarms to split into a main body and satellite swarm and thus violate the assumptions of the model.
In the case of a prolate axisymmetric ellipsoid b = c < a, we have K2 = K3 and from Eqs. (11)-(13) we get
K1
K3
= Z ∞
0
2,Z ∞
0
dv
dv
(1 + v)2 (p2 + v)
1
(1 + v) (p2 + v)
=
3
2
p(cid:16)p − p3 +pp2 − 1 cosh−1 p(cid:17)
2(cid:16)p2 − 1 − ppp2 − 1 cosh−1 p(cid:17) ,
(21)
TABLE I. Experimental data for the dependence of the mean swarm shape (given by the inertia eigenvalues I1, I2, I3) and
effective spring constants along the principle directions, on the swarm size given by the average radius Rs (binned averages).
9
Rs(mm) K1(1/sec2) K2(1/sec2) K3(1/sec2) η1 = I1/I2 η2 = I1/I3
43.22
52.58
58.40
63.97
70.77
76.53
81.59
88.15
98.62
108.89
120.99
15.83
12.33
10.62
10.12
9.09
8.83
7.36
6.65
5.94
6.94
4.82
14.72
11.06
9.50
10.06
7.98
7.60
6,55
6.11
5.49
6.34
4.62
11.80
8.34
7.17
5.83
4.84
5.33
3.31
3.10
3.05
4.56
3.15
1.38
1.27
1.37
1.26
1.32
1.25
1.40
1.26
1.43
1.64
1.57
1.81
1.70
1.68
1.61
1.77
1.54
1.98
1.62
1.74
2.07
2.06
where p ≡ a/c. For small deviations from spherical symmetry p = 1 + ǫ, ǫ ≪ 1 we have
K1
K3
= 1 +
6
5
ǫ + O(ǫ2).
(22)
This result is shown in Fig. 3b.
To summarize the results of this part, our AGM explains why the elongation of the swarms along the vertical axis
gives rise to lower effective spring constant that is observed in this direction, as well as to the different scaling with
the swarm size (Fig. 2).
B. The virial relation.
Despite the fact that adaptivity prevents us from formulating many conservation laws, we can still develop an
analog to the virial theorem, based on mass conservation. We can define analogous kinetic and potential energies,
and write a conservation law that relates them in case of stationarity. For this purpose we follow here Chandrasekhar
[42] and ideas that are used in galactic dynamics (see for instance [43]) to derive the virial equations for an ideal
self-gravitating fluid described in terms of density ρ(~r, t) and an isotropic pressure p(~r, t). The virial equations that
we derive here are equations of the second order, since they relate second-order moments. There are equations of
higher order as well. Later we will show how this derivation is related to the midge swarms.
We assume that, in addition to pressure gradients, the only fields that act on the fluid are the internal self-
gravitational field of the fluid ~g(in)(~r) and the external gravity field in the z direction, given by
Then the hydrodynamic equation governing the motion of the fluid is given by
~g(ext) = −g z.
where vi(~r, t) is the velocity of the fluid,
ρ
dvi
dt
= −
∂p
∂ri
+ ρ gi(~r),
d
dt
=
∂
∂t
+ vj
∂
∂rj
is the total time derivative (typically called the material derivative in fluid mechanics) and
gi(~r) = g(in)
i
(~r) − gδiz.
(23)
(24)
(25)
In order to obtain the second order virial equations we multiply Eq. (24) by rj and integrate over the entire volume
V . Integration of the term on the left hand side of Eq. (24) gives
ZV
ρ rj
dvi
dt
d3~r =ZV
ρ(cid:20) d
dt
(rj vi) − vj vi(cid:21) d3~r =
d
dt(cid:18)ZV
ρ vi rj d3~r(cid:19) − 2 Tij,
(26)
where we use integration by parts (assuming that the fluid is concentrated in a bounded region of space), and Π is
the total internal pressure of the fluid. The last term of Eq. (24) gives us a "potential energy" term:
Collecting the terms together we get
d
dt(cid:18)ZV
We can write Eq. (31) in a form similar to Newton's second law by introducing the moment of inertia tensor
ρ rjgi(~r)d3~r.
Wij =ZV
ρ vi rj d3~r(cid:19) = 2 Tij + Wij + δijΠ.
Iij =ZV
ρ ri rjd3~r,
10
(27)
(28)
(29)
(30)
(31)
(32)
(33)
where
is the kinetic energy tensor. The last equality of Eq. (26) is a direct consequence of mass conservation, namely
The same procedure with the first term on the right hand side of Eq. (24) gives
d
ρ vivj d3~r,
Tij =ZV
dt(cid:18)ZV
ρ(~r, t)d3~r(cid:19) = 0.
d3~r = δijZV
∂p
∂ri
p d3~r ≡ δijΠ,
ZV
rj
and since the tensors on the right hand side of Eq. (31) are symmetric we get
1
2
d2 Iij
dt2 = 2 Tij + Wij + δij Π.
When the system is stationary the moment of inertia does not change over time, and we arrive at the tensor form of
the virial theorem:
In order to apply the tensor virial equations presented above to the midge swarms, we will have to write its discrete
analogues for N particles with equal (unit) masses. The moment of inertia tensor is now
2 Tij + Wij + δijΠ = 0.
(34)
where the bar denotes an average value per midge. Its derivative with respect to time can give us an indication for
deviations of the system from stationarity, namely
¯I ij ≡
1
N
nrj
ri
n,
NXn=1
(35)
¯M ij ≡
d ¯I ij
dt
=
1
2 N
NXn=1(cid:0)ri
nvj
n + vi
nrj
n(cid:1) .
(36)
We use upper indices for the quantities that are defined with discrete summation. In Fig. 5 we show the values of
¯M ij taken for swarms of different sizes. Out of 126 measured swarms, we consider in this section binned data from
69 swarms that consisted of five midges or more, since for swarms with too few midges the average is meaningless.
In addition we take time averages of the quantities over roughly one minute, so that the swarm is approximately in
a steady state. The average values of the different components of ¯M ij are small compared to the typical angular
momentum, which is two orders of magnitude larger. We therefore conclude that the midge swarms are stationary,
and we therefore expect that the virial relation (Eq. (34)) should hold. Deviations from stationarity might occur
due to influx or outflux of midges (negligible) or irreversible processes. The small increase in the values of ¯M zz and
s
2
m
c
2
0
1
0.06
0.04
0.02
0.00
-0.02
-0.04
M xx
M yy
M zz
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae ae
ae ae
ae
ae
ae
ae
ae ae ae
ae ae ae
ae
ae
ae ae
ae ae
ae ae
ae ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
s
2
m
c
2
0
1
70
80
90 100 110 120
0.06
0.04
0.02
0.00
-0.02
-0.04
M xy
M xz
M yz
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae ae ae ae
ae ae ae ae
ae
ae
ae ae
ae ae
ae
ae
ae
ae
ae ae
ae ae
ae
ae
ae
ae
ae ae
ae ae
ae ae
ae ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
11
ae
ae
ae
ae
ae
ae
70
80
90 100 110 120
Rsmm
Rsmm
Figure 5. The six components of the tensor ¯M ij (Eq. (36)) that captures deviations from stationarity, as a function of the
swarm size Rs
¯M yz for large swarms might be an indication for such an irreversible process, such as fragmentation as a result of the
elongation along the vertical direction.
Given the stationarity that we have found in the swarms, we now test the validity of the virial relation (Eq. (34)),
using its discrete analogue. The mean kinetic energy tensor of a midge in the swarm is
and
¯T ij ≡
1
2 N
nvj
vi
n,
NXn=1
¯W ij ≡
1
2 N
NXn=1
ri
nF j
n + F i
nrj
n
is its "potential energy" tensor. The discrete virial equation is therefore
2 ¯T ij + ¯W ij + δij ¯Π = 0,
(37)
(38)
(39)
where ¯Π is the average pressure on a midge. This pressure term represents the effect of the mutual repulsion when
the midges arrive too close to each other. It is easier to interpret and check the trace of the tensorial equation (39),
namely the scalar form of the virial equation
2 ¯T + ¯W + 3 ¯Π = 0.
(40)
Here ¯T is the mean total kinetic energy of a midge, ¯W is its mean "potential energy", and ¯Π is the mean isotropic
pressure on a midge. In Fig. 6 we show the measured ¯T and − ¯W /2 for different swarm sizes, when integrating over
all the midges in the swarm. We can see that they are approximately constant as functions of Rs and their mean
values are
h ¯Ti = (3.42 ± 0.08) · 102
h− ¯W /2i = (2.80 ± 0.08) · 102
cm2/s2,
cm2/s2.
Therefore, according to Eq. (40), the difference between the two gives the mean pressure in the swarm:
¯Π = −(0.41 ± 0.07) · 102cm2/s2.
(41)
(42)
In order to confirm that the identification of the mean pressure is correct, we consider each diagonal component of
Eq. (39) separately as is shown in Fig. 7. From the mean values we get that
2 ¯T xx + ¯W xx = (0.42 ± 0.03) · 102cm2/s2,
2 ¯T yy + ¯W yy = (0.46 ± 0.04) · 102cm2/s2,
2 ¯T zz + ¯W zz = (0.38 ± 0.03) · 102cm2/s2.
(43)
12
ae
à
ae
ae
ae
à
ae
à
ae
ae
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
2
s
2
m
c
2
0
1
2
W
-
,
T
5
4
3
2
1
0
60
70
80
90
100
110
120
130
Rs mm
Figure 6. The mean kinetic energy ¯T (red) and half of the mean potential energy − ¯W /2 (blue) as a function of the swarm size
Rs (Eq. (40)).
Since these values are roughly equal it gives us a good confirmation for the isotropic origin of the pressure term in the
virial relation. Note that the observation that ¯W is independent of the swarm size when calculated over the observed
density profile [4] of the whole swarm is in agreement with a calculation done using regular gravity [44].
In addition we see from Fig. 7 that the mean values of kinetic and potential energies, which are related to the
movement in the z direction, are significantly lower than the x and y directions. This is a result of the external
gravitational force in this direction that enters the equations (see Eq. (25)). From the point of view of the midge,
it seems that it is more beneficial to respond to the effective pull of neighboring midges, rather than waste energy
moving up and down against gravity (which holds no information regarding location within the swarm).
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
à
ae
ae
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
2
s
2
m
c
2
0
1
2
x
x
W
-
,
x
x
T
2.0
1.5
1.0
0.5
0.0
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
2
s
2
m
c
2
0
1
2
y
y
W
-
,
y
y
T
2.0
1.5
1.0
0.5
0.0
2
s
2
m
c
2
0
1
2
z
z
W
-
,
z
z
T
2.0
1.5
1.0
0.5
0.0
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae ae
à à
ae ae
ae ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
ae
60
70
80
90
100
110
120
130
60
70
80
90
100
110
120
130
60
70
80
90
100
110
120
130
Rs mm
Rs mm
Rs mm
Figure 7. The diagonal components of the kinetic ¯T ii (red) and potential − ¯W ii/2 (blue) energy tensors.
The off-diagonal components of the tensors T ij and W ij are roughly null (compared to the diagonal ones) as we
show in Fig. 8. This is expected for a system without dissipation, which maintains stationarity. Note that on the
"microscopic" level of each midge, this system is obviously dissipative and out of equilibrium as the midge consumes
chemical energy to power its flight. However, we find that on the coarse-grained scale of equivalent particles and
forces, the system is effectively dissipation-less.
So far, we have not considered the adaptive nature of the interactions. Let us assume a uniform density spherical
swarm and first calculate the dependence on Rs without adaptivity. In order to calculate the behavior of the potential
energy with Rs, we consider again the continuous version of the mean potential energy
¯W =
3Xk=1Z ρ(~r) rk (Fad)k d3~r,Z ρ(~r)d3~r,
(44)
where rk is the k-th component of the vector ~r.
ae
ae
à
ae
à
ae
ae
ae
à
ae
ae
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
2
s
2
m
c
2
0
1
2
z
x
W
-
,
z
x
T
0.4
0.2
0.0
-0.2
-0.4
ae
ae
à
ae
à à
ae ae
ae
ae
ae
ae
à
ae
ae ae
ae
à à
ae
ae
ae
ae
ae
ae
ae
à
ae
à
ae
ae
ae
ae
ae
à
ae
ae
ae
à
ae
ae
ae
à
ae
2
s
2
m
c
2
0
1
2
z
y
W
-
,
z
y
T
0.4
0.2
0.0
-0.2
-0.4
ae
ae
à
ae
ae
ae
à
ae
ae
à à à à
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
13
ae
ae
à
ae
à
ae
ae
ae
ae
à
ae
ae
à
ae
ae
ae
à
ae
ae
ae
2
s
2
m
c
2
0
1
2
y
x
W
-
,
y
x
T
0.4
0.2
0.0
-0.2
-0.4
ae
ae
à
ae
ae
ae
à
ae
ae
à à
ae
ae
ae
ae
ae ae
ae
à
ae
60
70
80
90
100
110
120
60
70
80
90
100
110
120
60
70
80
90
100
110
120
Rs mm
Rs mm
Rs mm
Figure 8. The off-diagonal components of kinetic ¯T ij (red) and potential − ¯W ij/2 (blue) energy tensors.
In the case of a uniform spherical symmetric swarm we have
where
¯W = Z r (Fad)r d3~r(cid:30) V,
V =
4 π R3
s
3
is the volume of the spherical swarm. For a linear restoring force of the form
where K is a positive constant, we get
(Fad)r = −K r
¯W = −
3
5
K R2
s.
In the case of gravitational interaction without adaptivity (Eq. (A1)), the effective spring constant is
and then we get a quadratic dependence on Rs:
K =
4
3
π ρ C,
¯W = −
4
5
π ρ C R2
s.
The force with the adaptive correction is obtained by substituting Eq. (A2) into Eq. (3):
Fef f (r) = −
3 R2
ad [2 Rs r − (R2
4 C r2
Rs+r(cid:17)]
s − r2) ln(cid:16) Rs−r
.
Substitution into Eq. (45) gives:
¯W = −
4 C Rs
ad Z 1
0
R2
x5 dx
2 x + (x2 − 1) ln(cid:16) 1−x
1+x(cid:17) ∼ −0.3
C
R2
ad
Rs
where x ≡ r/Rs.
Thus, for adaptive gravity in the purely adaptive regime, the potential energy behaves as
¯W ∝ Rs.
(45)
(46)
(47)
(48)
(49)
(50)
(51)
(52)
We now compare this result to the potential energy contribution of the midges near the center of the swarm, where
we expect to find the strongest effect of adaptivity. When we include all the midges of the swarm, the adaptivity is
not significant since the midges with large radius (r > Rs) and low density dominate the contribution to the total
potential energy. For this purpose we calculated the potential and kinetic energies of the same swarms (more than
14
five midges) but this time the summation was carried out only up to an upper cutoff (Rs and Rs/2). Near the center
of the swarm the density is roughly constant and high, so that the adaptive calculation of Eqs. (44)-(52) should apply.
The results are presented in Fig. 9. The mean kinetic energy is similar to the previous one (Fig. 6), i.e. independent
of Rs, except for some under-sampling of the fastest midges: h ¯Ti(r<Rs) = (3.07 ± 0.17) · 102 cm2/s2. The potential
energy is not constant and it is increasing as a function of Rs. In Fig. 9b we show that as the center of the swarm is
approached (i.e. r is constrained to smaller values), the increase of ¯W with Rs approaches a linear behavior, as we
predict in Eq. (52). Note that this is very different from the quadratic behavior for regular gravity (Eq. (49)). This
observation therefore constitutes an additional independent and strong source of support for our adaptive-gravity
form of the interactions within the midge swarm.
The virial relation also allows us to estimate the effect of the swarm size on the mean distance of closest approach
between two midges, which was found to decrease for increasing swarm size. This relation is given in the Appendix
(Eqs. (A1)-(A8), Fig. A4).
2
s
2
m
c
2
0
1
2
W
-
,
T
5
4
3
2
1
0
ae
à
ae ae
à à
ae
à
ae
à
ae ae ae ae
ae ae ae ae
ae
ae
ae
à
ae
ae
ae ae
à à
ae
ae
ae
ae
ae ae ae ae
ae
ae
ae
ae
ae
à
ae
ae
ae
ae
à
ae
ae
ae
60
70
80
90
100 110 120 130
Rs mm
a
1.0
ae
ae
ae
ae
ae
ae
à à à à
à
2
W
-
g
o
L
0.5
0.0
-0.5
ì ì
4.2
4.3
ì
ì
ì
4.4
ae
à
ae
à
à
ae ae
à
ae
à
ì
ì
ì
ì
ì
4.5
4.6
4.7
4.8
LogRs
b
Figure 9. (a) The mean kinetic ¯T (red) and potential − ¯W /2 (blue, constrained for r < Rs/2, Rs from bottom to top) energies
as a function of Rs. (b) Log-Log plot for the potential energy in the center of the swarm with a linear fit, constrained for
r < Rs/2, Rs, ∞ (yellow, purple and blue respectively). The power-law slopes are (yellow and purple respectively): 1.16 ±
0.12, 0.71 ± 0.05.
C. Particle-based simulations.
So far, we have demonstrated that the mean-field predictions of the AGM are in good agreement with the exper-
imental results. To explore the AGM further, we performed molecular dynamics-type, agent-based simulations. To
focus on the effects of the proposed adaptive-gravity interactions, we did not include in the simulation any explicit
noise terms. Thus, the motion of each midge arises purely from their mutual interactions. However, to maintain
numerical stability and cohesion of the swarm, it was necessary to augment the basic AGM (Eq. (1)) in three ways.
First, we added a short-range repulsion between midges. This repulsion prevents the fragmentation of the swarm
into small groups (such as pairs or triplets) that can become effectively isolated from the rest of the swarm due to
adaptivity-induced screening (Eq. (A1), Fig. A5); additionally, we previously found experimental evidence for this
kind of short-range repulsion in real swarms [5]. With this repulsion, the effective force in Eq. (1) becomes
~F i
ef f = CXj brij
r2
ij
[1 − 2 exp(−(rij /Rr)2](cid:18)
R−2
R−2
ad
ad +Pk ~ri − ~rk−2(cid:19) .
(53)
To prevent runaway midges and to be physically realistic, we also imposed a maximum midge velocity vmax. If the
midge velocity exceeds this value, we re-scale it as
And finally, we added an overall confining force to prevent the swarm from drifting in space, as discussed above.
This force is significant only far from the swarm center, and is intended to model the attraction to ground-based
~vnew = ~vvmax/v.
(54)
12
Rmarker
(ri/Rmarker)11,
~F i
marker = −bri
15
(55)
visual features that localize natural swarms [31]. It is given by
where the marker size is set by Rmarker = 1.5Rad. We do not, however, impose any differences between the vertical
direction and the in-plane directions, and so our simulated swarms are statistically isotropic in space.
The initial conditions in the simulations are as follows. The spatial coordinates for each midge are expressed in
spherical coordinates r, θ, and ϕ, and are initially random variables uniformly chosen between [0,Rmarker], [0,π] and
[0,2π], respectively. The initial velocities are all zero. We then update the locations and velocities of each of the
particles in time by solving the (Newton's) equations of motion of the particles (that is, ~vi = ~F i
marker) using
Runge-Kutta integration. Note that, unlike in real life and unlike in most models, the midges in these simulations do
not have any intrinsic self-propelled motion; rather, their motion arises purely due to interactions, and can itself be
viewed as an emergent property of the swarm.
ef f + ~F i
A sample trajectory of a single midge from a simulated swarm of N = 50 midges, is shown in Fig. 10a, which
qualitatively resembled an observed trajectory (Fig. 10b). The simulated mean acceleration of a midge towards the
swarm center as a function of the distance from the center is shown in Fig. 10c. We recover the linear behavior near
the swarm center, as expected, while the forces saturate near the swarm edge due to the Gaussian density profile of
the swarm (Fig. 1b). The slopes near the center define the effective spring constants, and are found to decrease with
the number of midges, just as they do in the experiments. In Fig. 10d, we plot the distribution of midge accelerations,
and find that it displays the same qualitative features found in the experiments [4]: the distributions are close to
Gaussian for very small accelerations, but show heavy, exponential tails for large accelerations. And just as in the
experiments, we find that these distributions are largely independent of the swarm size.
We also compared the velocity distributions from the simulations (Fig. 10e) as a function of the swarm size, and
again found the same trend observed in the experiments [4]: as the swarm size increases, the velocity distributions also
develop a long exponential tail. This behavior can be quantified by calculating the excess kurtosis of the x-component
velocity distribution as a function of the swarm size (Fig. 10f), which follows the same qualitative behavior seen in
experiments [31]. For very small swarm sizes, the excess kurtosis is slightly negative (meaning that the tails of the
velocity distribution are slightly sub-Gaussian), and becomes positive for larger swarms.
Let us note that in these simulations we did not fit any parameters in an aim to reproduce the experimental
observations quantitatively. Rather, we focus here on exploring the qualitative features that arise due to the adaptive-
gravity interactions. It is quite satisfying that the distinctly non-Gaussian distributions of the accelerations (Fig. 10d)
and of the velocity (Fig. 10e,f) already appear within our simple model, as well as the overall dependence of various
features on swarm size. We anticipate that a more detailed model that includes, for example, the stochastic motion of
an individual midge in isolation [31] or the earth's gravitational field may be able to capture the mean-field behavior
of the swarms quantitatively as well.
V. DISCUSSION
We have presented here a new model of collective behavior in insect swarms that is based on the way that midges are
thought to sense their environment, i.e. through acoustic signals. The AGM we have constructed introduces features
that are not typically considered in models of collective motion, including long-range interactions and a sensitivity to
the global properties of the group (through adaptivity). As we have shown, these features combine to produce group
cohesion as a natural emergent property. Basic assumptions of the model, such as the precise relation between the
received sound and the force produced by the midge await future direct experimentation, by, for example, studying
external acoustic perturbations of swarms [35]. However, by comparing the predictions of the model with detailed
statistical data extracted from real insect swarms measured in the laboratory, we have demonstrated that our model
is able to capture not just the cohesion of swarms but also many of their many-body dynamical properties. The
excellent agreement between the AGM and the observed behavior of both the spatial profile of the average forces
within the swarm (Figs. 2,4) and the virial relation (Fig. 9), gives strong support to the model, and to its two main
long-range (gravity-like) 1/r2 interactions and an adaptive response that renormalizes the effective forces
features:
according to the local noise amplitude.
To conclude, this model opens the door for further tests of the large-scale behavior and stability of swarms.
Intriguingly, the success of the adaptive-gravity approach raises the question of whether some of the well known
phenomena that occur in self-gravitating systems, such as the Jeans instability or gravitational collapse [45, 46], can
occur in this biological system as well. This will be probed in future experiments, using acoustic perturbations of the
swarm [35]. From the more general physics point of view, we introduce here a new class of "adaptive" interactions,
which have novel physical features as well as apply to other biological systems -- for example in the context of chemical
16
Figure 10. Comparison of the adaptive-gravity simulations to experimental observations. (a) A sample simulated trajectory
from a swarm of N = 50 midges, as compared with (b) an observed trajectory in a swarm of N = 25 midges [31]. (c) Simulated
mean acceleration along the x-axis as a function of position for swarms of various sizes (N = 15, 25, 40, 90, from top to bottom).
The top right inset shows the simulated linear behavior near the swarm center, and the dependence of the slope on the swarm
size (different colors) is similar to the observed behavior shown in Fig. 1a. (d) Observed acceleration distribution (left panel,
[4]), compared to the simulation results (right panel). Both display highly non-Gaussian tails. (e) Observed x-component
velocity distribution (left panel, [4]), compared to the simulation results (right panel). Both show that small swarms have a
roughly Gaussian velocity distribution, but large swarms develop heavy tails. This tendency is quantified in f, where the excess
kurtosis is plotted as a function of swarm size, and both observations (left panel, [31]) and simulations (right panel) indicate a
negative excess kurtosis for small swarms and a positive kurtosis (due to the roughly exponential tails in the distributions) for
large swarms. All the simulations were carried out using Rad = 10, Rr = 0.3Rad, Rmarker = 1.5Rad, and vmax = 1.
sensing, chemotaxis-driven interactions between swarming cells (see for example [47]). This work opens a new class of
physical systems with adaptive interactions, whose properties may be very different from their non-adaptive analogs.
APPENDIX
Appendix A: Adaptive-gravity potential between two midges
Let us consider two interacting midges. In this case the effective force (Eq. (1)) felt by one midge due to a second
can be written as (taking C = 1)
~Fef f = r12
1
~r1 − ~r22A(~r1 − ~r2)
(A1)
where r12 = (~r1 − ~r2)/~r1 − ~r2 is the unit vector connecting the two midges and the adaptivity function is given by
A(~r1 − ~r2) =
R−2
ad
R−2
ad + ~r1 − ~r2−2
.
(A2)
17
Rad is a measure of the maximal distance between midges over which the adaptivity of the midge acoustic sensing
can function. Beyond this distance, the sound reaching the midge is so weak that the hearing sensitivity (that is,
the effective gain provided by the adaptivity) of the midge reaches its maximal value. See Fig. A1. Note that as
the midges come closer than Rad, the effective force approaches a constant, while in pure gravity the force between
two bodies increases without bound as they come together, leading to collapse. For r12 ≥ Rad the effective force
approaches the long-range gravity behavior of 1/r2
12.
We can integrate this force and calculate the effective midge-midge potential to be
Uef f (r12) =
1
Rad
(arctan [r12/Rad] − π/2) .
(A3)
See Fig. A2. Unlike in gravity, where the potential energy diverges when the two particles overlap, due to adaptivity
the potential approaches a finite value with a "cusp" shape, so that it is linear in r12. For r12 ≥ Rad the effective
potential approaches the long-range gravity behavior of 1/r12.
Figure A1. Magnitude of the effective force Fef f calculated according to Eqs. (A1) and (A2) using Rad = 10 (blue line). The
red line denotes pure gravity without adaptivity, i.e. for A = 1.
Figure A2. Effective potential Uef f calculated from Eq. (A3) using Rad = 10 (blue line). The red line denotes pure gravity
without adaptivity, i.e. for A = 1.
Appendix B: Explicit Calculation of the Effective Force on a Midge
Let us compute explicitly the adaptive-gravitational field at a point inside a spherical swarm with radius Rs and
uniform density ρ, according to Eq. (3). We will use cylindrical coordinates (r, z, ϕ) and calculate the field at (r = 0,
18
z
(r',z')
A
z0
r
Figure A3. The cylindrical coordinates (r, z) (and ϕ) that we use for the calculation of the effective gravitational field at a
point A in a uniform-density spherical swarm.
z = z0) without loss of generality (the point A in Fig. A3). The symmetry of the problem implies that the field is
along the z axis. The contribution of a point at (r′, z′) to the gravitational force at (0, z0) is
and the angle is
1
r′2 + (z′ − z0)2 ,
Hence the gravitational force at z0 is
cos α =
.
z′ − z0
pr′2 + (z′ − z0)2
dz′Z √R2
s−z ′2
r′dr′
0
z′ − z0
[r′2 + (z′ − z0)2]
3
2
Fef f,g(z0) = 2 π C ρZ Rs
−Rs
4 π ρ C
= −
3
which agrees with the result obtained from Gauss's law in Eq. (5). The normalization factor, which is given by the
summation of the absolute values of the contributions to the point (0, z0), reads
z0,
(A1)
Ntot(z0) = 2 π ρZ Rs
−Rs
= πρ[2Rs −
To leading order in z0, we have
dz′Z √R2
0
(R2
s − z2
0)
z0
s−z ′2
r′dr′
r′2 + (z′ − z0)2
Rs + z0(cid:19)].
ln(cid:18) Rs − z0
Ntot(z0) = 4πρRs + O(z2
0),
(A2)
(A3)
which confirms the 1/Rs leading behavior of the effective spring constant in the purely adaptive regime (Eq. (3)).
We find that the profile of Fef f (z) has the following behavior (Fig. 1a): The force is approximately linear in the
distance out to at least half of the swarm size, and only grows faster than linear close to the edge. Note that we have
assumed a uniform distribution here. Since the real distribution is close to Gaussian (see Fig. (2) of [4]) and decreases
significantly when r ∼ Rs, we do not expect to see such an increase in the force near the edges in real swarms.
Appendix C: Limit of cylindrical swarm
In the case of an elongated swarm in the form of a cylinder with radius R and length 2 L, the gravitational force at
z0 is
Fef f,g(z0) = 2 π C ρZ L
dz′Z R
= −4 π ρ C (cid:18)1 −
−L
0
r′dr′
z′ − z0
3
2
[r′2 + (z′ − z0)2]
√L2 + R2(cid:19) z0 + O(cid:18) z3
L3(cid:19) ,
L
0
(A1)
When we have a very long cylinder L ≫ R,
The normalization factor in this case is
Fef f,g(z0) ∼ −2 π C ρ
R2
L2 z0
Ntot(z0) = 2 π ρZ L
dz′Z R
= 4 R π ρ arctan(cid:18) L
−L
0
r′dr′
r′2 + (z′ − z0)2
R(cid:19) + 2 L π ρ ln(cid:18)1 +
R2
L2(cid:19) + O(cid:16) z0
L(cid:17) ,
and for a long cylinder
From Eqs. (A2) and (A4) we get the effective spring constant in the vertical direction as
Ntot(z0) ∼ 2 π2 ρ R.
Kz ∼
C R
π L2
The radius of a long cylindrical swarm is
Then in terms of the swarm radius
Rs = hri =
2 πR L
−L dzR R
0
2 π R2 L
√r2 + z2 r dr
L
2
∼
+ O(cid:0)L−1(cid:1)
Kz ∝ R−2
s .
19
(A2)
(A3)
(A4)
(A5)
(A6)
(A7)
Appendix D: Calculation of the effective spring constants in the ellipsoidal approximation
Here we show how to derive Eqs. (11)-(13). We start from a computation of the gravitational potential inside an
ellipsoid, and from it we will derive the effective "spring constants." The derivation here follows Ref. [48].
Let us start from an ellipsoid centered at the origin such that
x2
a2 +
y2
b2 +
z2
c2 = 1,
(A1)
whose semi-axes are a, b, and c, and where we assume c < b < a without loss of generality. Let us now consider a
family of ellipsoids
x2
a2 +
y2
b2 +
z2
c2 = u2
(A2)
for 0 < u < 1.
Our goal is to find the gravitational potential of an ellipsoid (A1) with a uniform mass density ρ. For this purpose,
we will calculate the potential due to an ellipsoidal shell between u and u + du and then sum over all shells for
0 < u < 1. It is convenient to introduce ellipsoidal coordinates (λ, µ, ν) defined by
x2 =
y2 =
z2 =
(a2 u2 + λ)(a2 u2 + µ)(a2 u2 + ν)
u4 (a2 − b2)(a2 − c2)
(b2 u2 + λ)(b2 u2 + µ)(b2 u2 + ν)
u4 (b2 − a2)(b2 − c2)
(c2 u2 + λ)(b2 u2 + µ)(b2 u2 + ν)
,
,
,
u4 (c2 − a2)(c2 − b2)
where the ellipsoidal coordinates (λ, µ, ν) have the following ranges:
λ > −c2 u2 > µ > −b2 u2 > ν > −a2u2.
(A3)
(A4)
(A5)
(A6)
Note that in the definition of ellipsoidal coordinates that we use, λ, µ, ν depend on u since we define them for the
ellipsoidal shell (A2). Surfaces of constant λ are ellipsoids:
x2
a2 u2 + λ
+
y2
b2 u2 + λ
+
z2
c2 u2 + λ
= 1.
(A7)
Thus, λ is the equivalent of the radial coordinate for ellipsoids. The original shell surface is λ(u) = 0 and λ(u) > 0
covers the region outside the shell (A2). A natural assumption (based on the ellipsoidal symmetry) is that the potential
depends only on λ and not on µ or ν. In this case, the Laplace equation outside the shell (for φ(λ, u), λ > 0) is
20
where
4pβ(λ, u)
(λ − µ)(λ − ν)
∂
∂λ(cid:18)pβ(λ, u)
∂φ
∂λ(cid:19) = 0,
β(λ, u) ≡ (a2 u2 + λ)(b2 u2 + λ)(c2 u2 + λ).
The solution of this equation can be written as
φ(λ, u) = −Z ∞
λ(u)
,
u2 D ds
pβ(s, u)
where D is a constant that will be related to the surface mass density of the shell, and the u2 factor is due to the fact
that the total charge scales as u2. Using Gauss's law for the shell,
(A8)
(A9)
(A10)
(A11)
(A12)
(A13)
(A14)
(A15)
(A16)
(A17)
(A18)
we get
− ∇φ · nλλ=0 = 4 π σ,
where gλλ comes from the metric in ellipsoidal coordinates:
= 4 π σ,
∂φ
−
1
√gλλ
∂λ(cid:12)(cid:12)(cid:12)(cid:12)λ=0
∂λ(cid:19)2
∂λ(cid:19)2
+(cid:18) ∂z
+(cid:18) ∂y
gλλ =(cid:18) ∂x
∂λ(cid:19)2
=
(λ − µ)(λ − ν)
4 β(λ, u)
,
so that
On the other hand, the surface mass density can be expressed as
σ = ρ dh,
D = 2 πσ√µν.
where dh is the local thickness of the shell. In order to find the thickness dh, let us start with a calculation of the
distance from the origin to a point (x, y, z) on the surface of the ellipsoid shell. The normal to the point (x, y, z) is
and the distance of the origin to a point (x, y, z) on the surface of the ellipsoid shell is then
a4 +
y2
b4 +
z2
c4 ,
z
a2 ,
y
b2 ,
n =(cid:16) x
c2(cid:17) /r x2
h = (x, y, z) · n = u2/r x2
a4 +
y2
b4 +
z2
c4 =
u a b c
√µν
.
Therefore the local thickness is
dh = du
a b c
√µν
,
which allows us to express the surface density σ of the shell in terms of ρ as
Substituting into the expression for D (Eq. (A14)), we get
σ = ρ
a b c
√µν
du.
The potential of such a shell is thus
D = 2 π a b c ρ du.
φ(λ, u)du = −2 π a b c ρ u2 duZ ∞
λ(u)
Then the gravitational potential of an ellipsoid is obtained by integration over such shells,
ds
.
pβ(s, u)
φ(λ, u)du.
0
U (λ) =Z 1
U (v) = −2 π a b c ρZ 1
0
uZ ∞
v(u)
du,
dt
pβ(t)
Let us change variables to v ≡ λ/u2, t ≡ s/u2, so that
where β(t) ≡ (a2 + t)(b2 + t)(c2 + t). Since we are interested in the force inside the ellipsoid, let us consider the
potential field at a point (x0, y0, z0) inside the ellipsoid. It corresponds to a shell at u0 (0 < u0 < 1):
x2
0
a2 +
y2
0
b2 +
z2
0
c2 = u2
0.
(A24)
For shells inside this shell (u < u0) the point (x0, y0, z0) is outside and therefore v = v(u) is the lower bound of the
integral. Shells outside this shell (u > u0) contribute the same value as on the shell surface (since u0 is inside the
shells) and then the lower bound is v = 0. Therefore
21
(A19)
(A20)
(A21)
(A22)
(A23)
(A25)
(A26)
(A28)
(A29)
(A30)
Integration by parts of the first integral gives
dt
pβ(t)!# .
dv
du
du.
U (v) = −2 π a b c ρ"Z u0
0
u Z ∞
Z u0
0
0
dt
dt
u0
0
dt
v(u)
v(u)
v(u)
u2
u du(cid:19) Z ∞
pβ(t)! du +(cid:18)Z 1
u Z ∞
pβ(t)#u0
pβ(t)! du =" u2
+Z u0
2 Z ∞
2pβ(v)
u Z ∞
pβ(t)! du =
2Z ∞
2 Z ∞
Z u0
pβ(t) −
pβ(v)! dv.
0 1 − u2
U (v) = −π a b c ρZ ∞
u2 dv
pβ(v)
u2
0
v(u)
dt
1
0
0
0
0
dt
Notice that v(u0) = 0 and v(0) = ∞. Hence,
Together with the second integral in (A25) we get
From (A7) we get the following relation between u and v:
Then the potential is
u2 =
x2
a2 + v
+
y2
b2 + v
+
z2
c2 + v
.
U (x, y, z) = −π a b c ρZ ∞
0 (cid:18)1 −
x2
a2 + v −
y2
b2 + v −
z2
c2 + v(cid:19)
dv
1
pβ(v)
.
(A27)
for a point (x, y, z) inside the ellipsoid. Effectively, we have a harmonic potential with different "spring constants" in
each direction. Taking the derivative of Eq. (A30) with respect to the coordinates x,y and z gives us Eqs. (11)-(13),
respectively.
To calculate the correction due to adaptivity to leading order, we have to write the expression for the total buzzing
noise (A2) for the ellipsoidal case. The only change is in the limits of integration and the fact that there is no longer
any cylindrical symmetry. Here we integrate over an ellipsoid and get
22
r′dr′
dz′Z Rel
0
0
0
Ntot(z0)=2 ρZ 2 π
dϕZ c
=2 √2 ρ a b cZ 2 π
0
r′2 + (z′ − z0)2
arccoth(cid:18) √2 a b
√Q(a,b,c,ϕ)(cid:19)
pQ(a, b, c, ϕ)
dϕ
(A31)
+ O(z2
0 ),
Q(a, b, c, ϕ) ≡ 2 a2 b2 − (a2 + b2)c2 + (a2 − b2)c2 cos 2 ϕ,
R2
el ≡
1 − z ′2
a2 + sin2 ϕ
b2
c2
cos2 ϕ
.
where
and
It turns out that this expression is symmetric under the interchange of a, b and c, and therefore there is no change to
the ratios of the effective spring constants due to the asymmetries of the ellipsoid in the leading-order correction.
Appendix E: Mean closest approach distance.
It was observed that the mean nearest-neighbor distance dnn between midges decreases with increasing swarm size
[4, 7, 31] (Fig. A4). We now demonstrate that this decrease can be explained to arise from the adaptive nature of
the interactions. Due to the adaptive interactions (Eq. (1)), when two midges happen by chance to come very close
to each other, they may form a bound pair that is effectively screened from the rest of the swarm. In order to study
the closest approach of midges in the swarm, let us consider the interaction of such a pair in the background provided
by the rest. We assume that the separation between the two midges is small compared to their distance to the rest
of the swarm, and thus that the interactions with the rest of the swarm will be negligible except for a contribution to
the adaptivity factor in Eq. (1), so that the effective force felt by one member of the pair is
~Fef f,pair ≃ C r12
1
~r1 − ~r22(cid:18)
ad + Ibackground + ~r1 − ~r2−2(cid:19) .
R−2
ad
R−2
(A1)
Ibackground is the sum over all the midges in the background. Near the center of the swarm this "background noise"
takes, according to Eq. (A3), the form
where we assume spherical symmetry (radius Rs) and a constant density ρ of the swarm.
Integrating this force we can calculate the effective two-body potential to be (see Eq. (A3))
Ibackground = 4 π ρ Rs,
(A2)
(A3)
(A4)
Upair =
where rpair ≡ ~r1 − ~r2 and
C
γ Rad(cid:18)arctan(cid:18) γ rpair
Rad (cid:19) −
π
2(cid:19) ,
γ ≡q1 + 4 π ρ R2
ad Rs.
We thus effectively have two-body motion under the influence of a mutual central force. Note that in the case of
two bodies, additivity of the effective force is valid and as a result we can use all the conservation laws of a central
force (namely energy and angular momentum). Therefore this two-body system can be reduced to an equivalent
one-dimensional motion in the effective potential
23
Uef f,12(rpair) =
β
r2
pair
+
C
γ Rad(cid:18)arctan(cid:18) γ rpair
Rad (cid:19) −
π
2(cid:19) ,
(A5)
where β is a positive constant [49] and the first term is the "centrifugal barrier." Let us assume that the system of
the two midges has an energy E. Then the minimal distance dnn is a turning point of the effective potential, given
by the solution of the equation
For very short distances rpair ≪ Rad, it is determined by the "centrifugal barrier", so that
Uef f,12(dnn) = E.
β
nn −
d2
π C
= E.
2 γ Rad
For swarms where Rs < √N Rad, we can expand dnn to obtain
2 γ Rad! 1
dnn =
E + π C
β
2
≃r β
E
.
(A6)
(A7)
(A8)
For the central part of spherical swarms the mean energy of a midge depends on the swarm size (Eq. (52)), as
E = ¯T + ¯W, and ¯W ∝ Rs (Fig. 9). We therefore expect the closest approach distance to decrease with increasing
swarm size. The experimental data shown in Fig. A4 for the center of the swarms (where our adaptivity calculation
is applicable), indicates a very weak decrease with swarm size, consistent with this prediction.
n
n
d
g
o
L
ae
aeaeae
aeae
ae
ae
ae
5.0
4.5
4.0
3.5
3.0
2.5
2.0
ae
ae ae
aeae
ae
ae
ae
aeae
ae ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
60 70 80 90 100 110 120
Rs
Figure A4. Measured mean minimal midge separation dnn in different swarms (with at least four midges) [31], plotted against
the mean swarm radius Rs (blue circles: raw data, red circles: binned averages). We observe a decrease in the average minimal
separation with increasing swarm size. The AGM predicts a shallow decrease, as indicated by the dashed line (Eq. (A8)), which
is in good agreement with the binned data (using the observed energies for ¯T , ¯W from Fig. 9, and treating β as the only fit
parameter).
Appendix F: Tendency for swarm fragmentation due to the adaptive "screening"
As discussed above, two nearby midges may form a bound pair due to screening by adaptivity. When the pair of
midges are very close to each other, we can have that
1
~ri − ~rnn2 ≫
1
R2
ad
+ Ibackground
(A1)
24
Figure A5. (a) Traces of a simulated swarm with N = 50 midges, demonstrating that without close-range repulsion there is a
tendency to form very tight pairs that leave the swarm. (b) As in (a) but with short-range repulsion that prevents the formation
of bound pairs, but no "marker" potential. (c) Traces of simulated trajectories of a swarm with N = 5 midges, demonstrating
the need for the "marker" potential and short-range repulsion to prevent midges from leaving the swarm.
which means that the interactions with all the other midges is reduced immensely by the huge factor of Eq. (1) in the
denominator. The resulting behavior is shown in Fig. A5. Including the "marker" potential and short-range repulsion
(Eqs. (53)-(55)) prevents midges from leaving the swarm (Fig. A5b,c), while otherwise pairs may form and shoot out
of the swarm (Fig. A5a,b). Occasionally, larger number of midges (such as triplets) also form such fragments.
[1] J. K. Parrish and L. Edelstein-Keshet, Science 284, 99 (1999).
[2] G. Gerisch, Annu. Rev. Physiol. 44, 535 (1982).
[3] K. Drescher, K. C. Leptos, I. Tuval, T. Ishikawa, T. J. Pedley, and R. E. Goldstein, Phys. Rev. Lett. 102, 168101 (2009).
[4] D. H. Kelley and N. T. Ouellette, Sci. Rep. 3, 1073 (2013).
[5] J. G. Puckett, D. H. Kelley, and N. T. Ouellette, Sci. Rep. 4, 4766 (2014).
[6] A. Attanasi, A. Cavagna, L. Del Castillo, I. Giardina, S. Melillo, L. Parisi, O. Pohl, B. Rossaro, E. Shen, E. Silvestri, and
M. Viale, PLoS Comput. Biol. 10, e1003697 (2014).
[7] A. Attanasi, A. Cavagna, L. Del Castillo, I. Giardina, S. Melillo, L. Parisi, O. Pohl, B. Rossaro, E. Shen, E. Silvestri, and
M. Viale, Phys. Rev. Lett. 113, 238102 (2014).
[8] M. Ballerini, N. Cabibbo, R. Candelier, A. Cavagna, E. Cisbani, I. Giardina, A. Orlandi, G. Parisi, A. Procaccini, M. Viale,
and V. Zdravkovic, Anim. Behav. 76, 201 (2008).
[9] M. Nagy, Z. ´Akos, D. Biro, and T. Vicsek, Nature 464, 890 (2010).
[10] Y. Katz, K. Tunstrom, C. C. Ioannou, C. Huepe, and I. D. Couzin, Proc. Natl. Acad. Sci. USA 108, 18720 (2011).
25
[11] J. E. Herbert-Read, A. Perna, R. P. Mann, T. M. Schaerf, D. J. T. Sumpter, and A. J. W. Ward, Proc. Natl. Acad. Sci.
USA 108, 18726 (2011).
[12] A. Berdahl, C. J. Torney, C. C. Ioannou, J. J. Faria, and I. D. Couzin, Science 339, 574 (2013).
[13] W. D. Hamilton, J. Theor. Biol. 31, 295 (1971).
[14] J. A. Downes, Annu. Rev. Entomol. 14, 271 (1969).
[15] S. J. Portugal, T. Y. Hubel, J. Fritz, S. Heese, D. Trobe, B. Voelkl, S. Hailes, A. M. Wilson, and J. R. Usherwood, Nature
505, 399 (2014).
[16] A. Okubo, Adv. Biophys. 22, 1 (1986).
[17] A. J. Bernoff and C. M. Topaz, SIAM Review 55, 709 (2013).
[18] T. Vicsek, A. Czir´ok, E. Ben-Jacob, I. Cohen, and O. Shochet, Phys. Rev. Lett. 75, 1226 (1995).
[19] I. D. Couzin, J. Krause, R. James, G. D. Ruxton, and N. R. Franks, J. Theor. Biol. 218, 1 (2002).
[20] N. T. Ouellette, Pramana-J. Phys. 84, 353 (2015).
[21] N. W. F. Bode, A. J. Wood, and D. W. Franks, Behav. Ecol. Sociobiol. 65, 117 (2011).
[22] M. Ballerini, N. Cabibbo, R. Candelier, A. Cavagna, E. Cisbani, I. Giardina, V. Lecomte, A. Orlandi, G. Parisi, A. Pro-
caccini, M. Viale, and V. Zdravkovic, Proc. Natl. Acad. Sci. USA 105, 1232 (2008).
[23] F. Ginelli and H. Chat´e, Phys. Rev. Lett. 105, 168103 (2010).
[24] J. Buhl, D. J. T. Sumpter, I. D. Couzin, J. J. Hale, E. Despland, E. R. Miller, and S. J. Simpson, Science 312, 1402
(2006).
[25] R. Lukeman, Y.-X. Li, and L. Edelstein-Keshet, Proc. Natl. Acad. Sci. USA 107, 12576 (2010).
[26] S. J. Guy, S. Curtis, M. C. Lin, and D. Manocha, Phys. Rev. E 85, 016110 (2012).
[27] H. Chat´e and M. A. Munoz, Physics 7, 120 (2014).
[28] M. V. Federova and R. D. Zhantiev, Entomol. Rev. 89, 896 (2009).
[29] J. Sueur, E. J. Tuck, and D. Robert, The Journal of the Acoustical Society of America 118, 530 (2005).
[30] D. J. G. Pearce, A. M. Miller, G. Rowlands, and M. S. Turner, Proc. Natl. Acad. Sci. USA 111, 10422 (2014).
[31] J. G. Puckett and N. T. Ouellette, Journal of The Royal Society Interface 11, 20140710 (2014).
[32] R. Y. Tsai, IEEE J. Robotic. Autom. RA-3, 323 (1987).
[33] N. T. Ouellette, H. Xu, and E. Bodenschatz, Exp. Fluids 40, 301 (2006).
[34] H. Xu, Meas. Sci. Technol. 19, 075105 (2008).
[35] R. Ni, J. G. Puckett, E. R. Dufresne, and N. T. Ouellette, Phys. Rev. Lett. 115, 118104 (2015).
[36] J. G. Puckett, R. Ni, and N. T. Ouellette, Phys. Rev. Lett. 114, 258103 (2015).
[37] M. Lighthill, in Proc. Roy. Soc. A, Vol. 267 (1962) pp. 147 -- 182.
[38] H. Bennet-Clark, Philosophical Transactions of the Royal Society B: Biological Sciences 353, 407 (1998).
[39] O. Shoval, L. Goentoro, Y. Hart, A. Mayo, E. Sontag, and U. Alon, Proc. Natl. Acad. Sci. USA 107, 15995 (2010).
[40] L. J. Cator, K. R. Ng'Habi, R. R. Hoy, and L. C. Harrington, Behav. Ecol. 21, 1033 (2010).
[41] W. Greiner, Classical mechanics: point particles and relativity (Springer, 2004).
[42] S. Chandrasekhar, Ellipsoidal figures of equilibrium, Silliman Foundati Lect. (Yale Univ. Press, New Haven, CT, 1969).
[43] J. Binney and S. Tremaine, Galactic dynamics (Princeton university press, 2011).
[44] D. Gorbonos, R. Ni, J. G. Puckett, N. T. Ouellette, and N. Gov, work in progress (2015).
[45] J. Jeans, Monthly Notices of the Royal Astronomical Society 82, 122 (1922).
[46] J. Binney, Annual Review of Astronomy and Astrophysics 20, 399 (1982).
[47] T. Lammermann, P. V. Afonso, B. R. Angermann, J. M. Wang, W. Kastenmuller, C. A. Parent, and R. N. Germain,
Nature 498, 371 (2013).
[48] Based on: Wei Cai, "Appendix - Potential Field of a Uniformly Charged Ellipsoid," as part of the lecture notes to the
course: "Theory and Applications of Elasticity," Department of Mechanical Engineering, Stanford University, CA 94305-
4040. http://micro.stanford.edu/∼caiwei/me340/A Ellipsoid Potential.pdf.
[49] In the case of gravity we take β = l2
2 µ where l is the angular momentum and µ is the reduced mass of the system.
Acknowledgements. N.S.G. and D.G. thank Sam Safran for useful discussion. J.G.P., R.N., and N.T.O. acknowl-
edge support from the US Army Research Office under grant W911NF-13-1-0426. N.S.G. gratefully acknowledges
funding from the ISF (Grant No. 580/12). This research is made possible in part by the generosity of the Harold
Perlman Family.
|
1011.3051 | 1 | 1011 | 2010-11-12T21:10:36 | Life around the scallop theorem | [
"physics.bio-ph",
"cond-mat.soft",
"physics.flu-dyn"
] | Locomotion on small scales is dominated by the effects of viscous forces and, as a result, is subject to strong physical and mathematical constraints. Following Purcell's statement of the scallop theorem which delimitates the types of swimmer designs which are not effective on small scales, we review the different ways the constraints of the theorem can be escaped for locomotion purposes. | physics.bio-ph | physics |
Life around the scallop theorem
Department of Mechanical and Aerospace Engineering,
Eric Lauga∗
University of California San Diego, 9500 Gilman Drive, La Jolla CA 92093-0411, USA.
(Dated: May 28, 2018)
Locomotion on small scales is dominated by the effects of viscous forces and, as a result, is
subject to strong physical and mathematical constraints. Following Purcell's statement of the scallop
theorem which delimitates the types of swimmer designs which are not effective on small scales, we
review the different ways the constraints of the theorem can be escaped for locomotion purposes.
I.
INTRODUCTION
Swimming cells, such as bacteria (prokaryotes) or spermatozoa (eukaryotes), represent the prototypical example of
active soft matter. They are active as they transform chemical energy (ATP for eukaryotes, ion flux for prokaryotes)
into mechanical work [1] and, as a result, are able to continuously change shape and move in viscous environments
[2]. As mechanical entities, cells belong to the world of soft matter, displaying complex rheological properties on a
range of time and spatial scales and responding to external forcing in a time-dependent and nonlinear fashion [3].
In their micron-size environment, the fluid forces acting on swimming cells are dominated by the effect of viscous
dissipation [4, 5]. Seminal papers in the 1950s laid the ground work for detailed investigations on the hydrodynamics
of cell locomotion [6–9], with the main goal of predicting cell kinematics, energetics, the interactions with their
environment, and the general importance of fluid forces in biological form and function [10–15].
In 1977, Purcell's influential paper "Life at low Reynolds number" put a somewhat different spin on a field which
was already mature [16]. In it, Purcell brought to light the counter-intuitive physical and mathematical constraints
arising from locomotion in an inertialess world. He demonstrated that for organisms moving in very viscous fluids,
there exists a class of shape change that can never be used for locomotion, a result beautifully summarized under
the name "scallop theorem", borrowing the name of such an organism - a hypothetical microscopic scallop - which
could not locomote in the absence of inertia.
In this short review, we look back at the scallop theorem, and pose the question: What are the basic ingredients
necessary to design swimmers able to move on small scales? What are the different ways offered by physics to get
around the constraints of the theorem? After stating the various assumptions for the theorem to be valid (§II), we
show how non-reciprocal shape changes (§III), inertia (§IV), hydrodynamic interactions (§V), and coupling with the
physical environment (§VI) can all be exploited to provide locomotion on small scales.
II. THE SCALLOP THEOREM
The scallop theorem has a relatively simple statement [16]. Consider a body changing shape in a time-periodic
fashion. In the absence of inertia, the equations describing the motion of an incompressible Newtonian fluid are Stokes
equations, which are linear and independent of time [4, 5]. In addition, in the absence of inertia, the swimmer remains
perpetually force- and torque-free [15]. Purcell's scallop theorem can then be stated as follows. If the sequence of
shapes displayed by the swimmer is identical to the sequence of shapes displayed when seen in reverse - so-called
reciprocal motion - then the average position of the body cannot change over one period. Another manner to describe
reciprocal motion is stated in Purcell's original paper as:
∗Electronic address: [email protected]
2
"... I change my body into a certain shape and then I go back to the original shape by going through the
sequence in revere... So, if the animal tries to swim by a reciprocal motion, it can't go anywhere."
Time is not explicitly mentioned in the theorem and in fact, because of the linearity and time-independence of the
equations, the rate at which the sequence of shapes is being displayed is irrelevant [15]. In Purcell's own words,
"Fast, or slow, it exactly retraces its trajectory, and it's back where it started."
Physically, the absence of time in the equations of motion means there is no intrinsic time scale to the swimming
problem, which prevents distinguishing between forward and backward in a reciprocal motion.
Purcell's statements may appear simple, but are in fact far-reaching. They form the basis of a purely geometrical
approach to cell locomotion [17–20] and have sparked considerable attention in the area of biolocomotion from the
physics and soft matter community - so much so that "Life at low Reynolds number" is now the most cited paper
in the field.
The name for the theorem originates from the simplest kind of reciprocal swimmers, namely those deforming with
a single degree of freedom, such as the hinge of a hypothetical micron-scale scallop. For any swimmer with a single
geometrical degree of freedom, say θ(t), then by properties of Stokes equations, its swimming speed, u, necessarily
scales as u ∼ θF (θ), which is always an exact derivative, and thus averages in time to zero hui = 0. Swimmers with
only one degree of freedom can thus never swim on small scales.
Strictly speaking, the scallop theorem is valid only with the following assumptions: a single swimmer displaying
reciprocal motion in an infinite quiescent Newtonian fluid and in the absence of inertia and external body forces1.
Examining each of these assumptions in detail suggests a way around the theorem and a design for a swimmer, which
we now review. As in Purcell's original paper, we will focus only on swimming by shape change or motion and we
will thus not consider chemical swimmers [21–24] or solid bodies powered by external fields [25].
III. NON-RECIPROCAL KINEMATICS
A. Biological swimmers: waves
The main message of Purcell's paper is that swimmers should change their shapes in a non-reciprocal fashion. The
manner in which motion occurs should thus indicate a clear direction of time, which leads naturally to the occurrence
of waves.
Indeed, most of swimmings cells locomote by using traveling wave-like deformation of their bodies or
appendages [10–15]. Swimming bacteria rotate one or more helical flagella using rotary motors embedded in the
cell walls [38–40] leading to flagella kinematics akin to that of traveling helical waves, and thus propulsion [41–43]
(Fig. 1a). Other types of bacteria swim using whole-body wave deformation propelled by flagella beneath the cell's
outer membrane [44] or wave-like propagation of kinks in their shapes in the absence of flagella [45]. Spermatozoa
and other singly flagellated eukaryotes swim using traveling waves [12] induced by molecular motors-driven internal
sliding of polymeric filaments inside the flagellum [46–48] (Fig. 1b). The flagella kinematics can be planar [49], helical
[43], or even doubly helical [50]. The many cilia covering some eukaryotes [51] also deform as so-called metachronal
waves [52–54] (Fig. 1c).
B. Synthetic swimmers
Beyond the swimming methods displayed by biological swimmers, some simpler modes of non-reciprocal motion
can be devised theoretically and in the lab.
1 The theorem is also valid in a confinement environment as long as the boundaries display no motion.
3
FIG. 1: Illustration of experimental and computational escapes from the scallop theorem. (a): E. coli bacterium with four
helical flagella [26]; (b): Superimposed pictures of a swimming spermatozoon of Ciona intestinalis [27]; (c): Paramecium cell
covered with short cilia [28]; (d): Macro-scale experimental realization of Purcell's three link swimmer [29]; (e): Micro-scale
experimental realization of three-sphere swimmer using optical tweezers [30]; (f): Computations showing the locomotion of
shape-changing vesicles [31]; (g): Microscopic realization of flexible swimming using elastic superparamagnetic filaments driven
by an external magnetic field (the two images show the filament deforming at different times) [32]; (h): Locomotion of flexible
Au/Ag/Ni nanowires swimmers driven by an external magnetic field [33]; (i): Experimental demonstration that a symmetric
flapping wing can undergo unidirectional locomotion if the Reynolds number is above a critical value [34]; (j): Computations
showing that a flexible, flapping wing with asymmetric actuation undergoes locomotion at all finite Reynolds number [35];
(k): Experimental measurement of the net flow and vorticity induced by a reciprocal flapper beneath a free surface [36];
(l): Computations for the net flow and vorticity induced by a small-amplitude reciprocal flapper in a polymeric fluid [37]. All
images reproduced with permission; (a): from Turner, Ryu, and Berg, J. Bacteriol. 182, 2793 (2000), Copyright 2000 American
Society for Microbiology; (b): from C. J. Brokaw, J. Exp. Biol. 43, 155 (1965), Copyright 1965 The Company of Biologists;
(c): Copyright CNRS Phototh`eque / Anne Aubusson-Fleury; (e): from Leoni et al. Soft Matt. 5, 472 (2009), Copyright 2009
Royal Society of Chemistry; (f): from Evans, Spagnolie, and Lauga, Soft Matter 6, 1737 (2010), Copyright 2010 Royal Society
of Chemistry; (g): from Dreyfus et al. Nature 437,862 (2005), Copyright 2005 Nature Publishing Group; (h): from Gao et al.
J. Am. Chem. Soc., 132, 14403 (2010), Copyright 2010 American Chemical Society; (i): from Vandenberghe, Childress, and
Zhang, Phys. Fluids 18, 014102 (2006), Copyright 2010 American Institute of Physics; (j): from Spagnolie et al., Phys. Fluids
22, 041903 (2010), Copyright 2010 American Institute of Physics; (k): from Trouilloud et al., Phys. Rev. Lett. 101, 048102
(2008), Copyright 2008 American Physical Society; (l): from Pak, Normand, and Lauga, Phys. Rev. E 81, 036312 (2010),
Copyright 2010 American Physical Society.
1.
Imposing non-reciprocal kinematics
As shown by Purcell, swimmers with a single degree of freedom cannot move. One needs therefore at least two
degrees of freedom and their prescribed variation in time should sweep a finite area in parameter space. In his original
paper, Purcell proposed such a swimmer [16], namely an elongated body with two rotational hinges [55–57] (Fig. 1d).
Subsequently, non-reciprocal swimmers of very simple shapes have been devised theoretically, including ones composed
of three spheres [30, 58–61] (Fig. 1e), two volume-changing spheres [62], and two-orientation changing spheres [63]
or ellispoids [64]. Beyond geometry, the two degrees of freedom could also be physical parameters, for example the
volume and spontaneous curvature of a lipid vesicle [31] (Fig. 1f). Alternatively, the swimmer's shape and deformation
change could be topologically equivalent to the inside-out rotation of a torus [65, 66] or tank-treading [67] for which
periodicity is achieved by a continuous series of displacements tangent to the swimmer shape. Continuous normal
flows in the form of fluid jets can also be used [68].
4
2. Flexible swimmers: Non-reciprocal kinematics from reciprocal forcing
A second class of simple swimmers can be designed for which a reciprocal actuation combined with flexibility or
elasticity can lead to kinematics of shape change which are non-reciprocal, and thus to locomotion.
The prototypical example of this class of swimmers is a flexible filament actuated periodically up and down at one
end where it is clamped, and free on the other [69]. If the filament is rigid, its motion is reciprocal and cannot be used
for propulsion. In contrast, if the filament is flexible and is actuated near the typical frequency at which viscous drag
and elastic forces balance, its shape as it is actuated up (respectively down) is concave (respectively convex), leading
to non-reciprocal kinematics and propagation of an elasto-hydrodynamic wave. Mathematically, the scallop theorem
breaks down because time enters the problem through the viscous drag term in the equation for the filament shape
(via a partial time-derivative), and thus a relevant time scale can be defined.
The generation of propulsive force and locomotion using flexibility filaments has been the center of many theo-
retical and computational investigations [69–72]. A macro-scale experiment confirmed the physical picture outlined
above [73]. Related phenomena include elastic buckling instabilities [74–76] and shape transitions [77–79] for rotated
elastic filaments. At the micro-scale, an experimental realization of a flexible swimmer was achieved using elastic
superparamagnetic filaments [80] actuated by external magnetic fields and attached to a red blood cell (Fig. 1g) [32],
prompting subsequent modeling efforts [81–84]. A similar implementation was achieved using a nanometric silver
filament attached to an externally-driven ferromagnetic nickel head (Fig. 1h) [33]. In all these cases however, it is
the presence of external torques (via external magnetic fields) that allows locomotion, and thus they do not represent
true self-propelled motion.
IV.
INERTIA
For the scallop theorem to be valid, all inertial terms in the equation of motion of the swimmer should be set to zero.
Naturally, they cannot exactly disappear unless no motion occurs, and thus a fundamental question arises, namely
how much inertia is needed to escape the constraints of the theorem? Is the scallop theorem valid only asymptotically,
or does it stand as long as inertia is below a certain limit? These questions were first posed by Dudley and Childress
[85] who studied the behavior of a mollusk able to use both reciprocal and non-reciprocal modes of locomotion, and
who postulated that a finite amount of inertia was necessary for locomotion to be able to occur.
Mathematically, three qualitatively different Reynolds numbers can be defined. Consider a swimmer of typical size
L and density ρs undergoing reciprocal motion of amplitude A and frequency ω in a Newtonian fluid of density ρ and
shear viscosity µ. Using a typical velocity scale U ∼ Aω, the natural Reynolds number for the reciprocal motion is
given by Re = ρLAω/µ, and is the one corresponding to the nonlinear advection term in the Navier Stokes equations
(for example, in water, Re ≈ 10−4 for E. coli while Re ≈ 10−2 for human spermatozoa). The oscillatory Reynolds
number, corresponding to the linear unsteady Stokes term, is given by Reω = ρL2ω/µ. Finally, the Reynolds number
based on the body inertia is Res = ρsL2ω/µ, sometimes called a Stokes number, which quantifies the typical ratio
between the rate of change of the swimmer momentum and the magnitude of the viscous forces in the fluid.
For small amount of inertia, the breakdown of the scallop theorem occurs either continuously or discontinuously
with these Reynolds numbers depending on the geometrical symmetries in the reciprocal actuation. In the case of
symmetric shapes - typically simple flappers - experiments and modeling demonstrated that a finite, order one,
amount of inertia is necessary, indicating a discontinuous transition through an inertial hydrodynamic instability
[34, 85–89] (Fig. 1i). As a difference, in the case of asymmetric shapes or actuation, the transition is continuous, with
locomotion occurring either as some power of Re [90] or both Re and Reω (with Re/Reω constant) [35, 91] (Fig. 1j).
Interestingly, for asymmetric shapes, a continuous transition with swimmer inertia was obtained in the absence of
fluid inertia (Re = Reω = 0), with locomotion occurring as powers of Res [92].
5
V. HYDRODYNAMIC INTERACTIONS
The inertialess scallop envisioned by Purcell as the prototypical non-swimmer is isolated in the fluid. It turns out
however that hydrodynamic interactions with other such non-swimmers, or more generally flexible entities, can be
exploited to swim. Physically, as cells or other synthetic swimming devices do work on the surrounding fluid, they
act as hydrodynamic disturbances on the otherwise-quiescent environment, thereby setting up flow fields which are
in general dipolar [15]. In biology these flow fields have important consequence on the generation of collective modes
of locomotion [93–97] and rheology at the whole-population level [98, 99].
Although a body undergoing reciprocal motion cannot swim, two bodies undergoing reciprocal motion with nontriv-
ial phase differences are able to take advantage of the unsteady hydrodynamic flows they create to undergo nonzero
collective and relative dynamics; there is thus no many-scallop theorem [100, 101]. As each reciprocal swimmer be-
haves in general as an unsteady dipole, the collective effect arises from the time-rectification of such unsteadiness,
and thus decays generically as 1/d3, where d is the typical swimmer-swimmer distance (or even faster if additional
geometrical symmetries are present [100, 102]). Naturally, two reciprocal non-swimmers taken as a whole are not
unlike a single non-reciprocal swimmer, although the qualitative details of their locomotion do differ [100].
Experimentally, this effect was demonstrated for hydrodynamic interactions between a rigid flapper, beating in a
reciprocal fashion, and a flexible boundary (free surface). The rectification of the reciprocal flow by the free surface
motion leads to flow and forces scaling quadratically with the applied flapping frequency, and the creation of a
reciprocal pump [36] (Fig. 1k). The experimental application of these ideas to a collection of free-swimming bodies
remains however to be confirmed. To generate reciprocal motion with nontrivial phase-differences, one possibility
would be to use elastic field-responsive particles under a uniform AC forcing; particles with different relaxation times
would respond to fields with different phases, and thus would be able to move collectively [100]. In the case of purely
identical non-swimmers, two of them cannot swim, but three or more are able to move [102]. In that case, the phase
differences in body kinematics are induced by hydrodynamic flows, leading to a slow 1/d7 effect [102].
VI. PHYSICAL ENVIRONMENT
In the scallop theorem, the assumption that locomotion takes place in a Newtonian environment is crucial, as it
allows the inertialess equations of fluid motion to be linear and independent of time. A change of the mechanical and
rheological properties of the fluid would however naturally lead to a different type of conclusion. Complex fluids are
abundant in biology, and cell locomotion often takes place in strongly elastic polymeric fluids [103–105], which has
been the focus of much recent work [106–111].
As the fluid becomes non-Newtonian, three different physical effects can potentially be exploited to generate small-
scale locomotion [112, 113]. First, complex fluids possess in general rheological properties which are rate dependent.
In particular, viscosities often display shear-thinning behavior, meaning they decrease with shear rates. In this type
of fluid, and in contrast with the Newtonian case, the rate at which the reciprocal sequence of shapes is being
displayed would matter, a result which could be used to design a reciprocal swimmer. This was recently demonstrated
theoretically for bodies swimming using a reciprocal helical actuation at different rates in model polymeric fluids [109].
The second physical effect to be exploited is that of normal stress differences, which arise from the stretching by
the flow of the microstructure suspended in the complex fluids. Normal stress differences scale quadratically with the
applied shear [112] and remain thus identical under a reversal a time, allowing propulsion. Locomotion using normal
stress differences was demonstrated theoretically for a three-dimensional body undergoing small-amplitude reciprocal
motion at constant rate [110]. The generation of forces and flow by reciprocal flapping was also reported [37, 114]
(Fig. 1l).
The last physical effect to be exploited is that of stress relaxation. Even for small-amplitude motion and linearized
dynamics, the simplest evolution equation for the stress in a polymeric fluid contains a memory term in the form of
a partial time derivative times a relaxation time. Whether, even in the linear regime, stress relaxation can be taken
advantage of for locomotion purposes is an intriguing, but yet unexplored, possibility.
6
VII. CONCLUSION
In this short review, we have used Purcell's scallop theorem as a framework to lay out the basic physical principles
behind the design of small-scale swimming devices. We have shown how non-reciprocal kinematics, inertia, hydrody-
namic interactions, and the nature of surrounding environment can all be physically exploited to achieve small-scale
propulsion. With advances in micro- and nano-fabrication, the discussion on the theorem can now move from that
akin to a mathematical exercise to a true engineering challenge.
As briefly mentioned in Ref. [13], there exists at least another class of body motion which always leads to zero
locomotion in a Newtonian fluid, namely those for which the time-reversal of the motion is identical to its mirror-
image (for example, the motion of a rod sweeping the envelope of a cone). The formal derivation of the complete class
of non-swimming body kinematics would provide a new thrust in small-scale locomotion research by allowing novel
opportunities to get around these mathematical constraints.
Acknowledgements
Discussions with Denis Bartolo are gratefully acknowledged. I thank B. Chan, A. E. Hosoi, A. Aubusson-Fleury
and the CNRS Phototh`eque for providing me with the images reproduced in Fig. 1c and d, as well as the authors
from Refs. [26, 27, 29–37] who gave us permission to reproduce their images. This work was supported in part by the
US National Science Foundation through grant number CBET-0746285.
[1] D. Bray, Cell Movements (Garland Publishing, New York, NY, 2000).
[2] G. T. Yates, Am. Sci. 74, 358 (1986).
[3] B. D. Hoffman and J. C. Crocker, Annu. Rev. Biomed. Eng. 11, 259 (2009).
[4] J. Happel and H. Brenner, Low Reynolds Number Hydrodynamics (Prentice Hall, Englewood Cliffs, NJ, 1965).
[5] S. Kim and J. S. Karilla, Microhydrodynamics: Principles and Selected Applications. (Butterworth-Heinemann, Boston,
MA, 1991).
[6] G. I. Taylor, Proc. Roy. Soc. A 209, 447 (1951).
[7] G. I. Taylor, Proc. Roy. Soc. A 211, 225 (1952).
[8] G. J. Hancock, Proc. Roy. Soc. Lond. A 217, 96 (1953).
[9] J. Gray and G. J. Hancock, J. Exp. Biol. 32, 802 (1955).
[10] J. Lighthill, Mathematical Biofluiddynamics (SIAM, Philadelphia, 1975).
[11] J. Lighthill, SIAM Rev. 18, 161 (1976).
[12] C. Brennen and H. Winet, Ann. Rev. Fluid Mech. 9, 339 (1977).
[13] S. Childress, Mechanics of Swimming and Flying (Cambridge University Press, Cambridge U.K., 1981).
[14] L. J. Fauci and R. Dillon, Ann. Rev. Fluid Mech. 38, 371 (2006).
[15] E. Lauga and T. R. Powers, Rep. Prog. Phys. 72, 096601 (2009).
[16] E. M. Purcell, Am. J. Phys. 45, 3 (1977).
[17] A. Shapere and F. Wilczek, Phys. Rev. Lett. 58, 2051 (1987).
[18] A. Shapere and F. Wilczek, J. Fluid Mech. 198, 557 (1989).
[19] J. Koiller, K. Ehlers, and R. Montgomery, J. Nonlinear Sci. 6, 507 (1996).
[20] E. Yariv, J. Fluid Mech. 550, 139 (2006).
[21] W. F. Paxton, K. C. Kistler, C. C. Olmeda, A. Sen, S. K. S. Angelo, Y. Y. Cao, T. E. Mallouk, P. E. Lammert, and
V. H. Crespi, J. Am. Chem. Soc. 126, 13424 (2004).
[22] R. Golestanian, T. B. Liverpool, and A. Ajdari, Phys. Rev. Lett. 94, 220801 (2005).
[23] J. R. Howse, R. A. L. Jones, A. J. Ryan, T. Gough, R. Vafabakhsh, and R. Golestanian, Phys. Rev. Lett. 99, 048102
7
(2007).
[24] G. Ruckner and R. Kapral, Phys. Rev. Lett. 98, 150603 (2007).
[25] S. J. Ebbens and J. R. Howse, Soft Matt. 6, 726 (2010).
[26] L. Turner, W. S. Ryu, and H. C. Berg, J. Bacteriol. 182, 2793 (2000).
[27] C. J. Brokaw, J. Exp. Biol. 43, 155 (1965).
[28] CNRS phototh`eque, URL http://phototheque.cnrs.fr/.
[29] B. Chan, Ph.D. thesis, Massachusetts Institute of Technology, Cambridge, MA (2009).
[30] M. Leoni, J. Kotar, B. Bassetti, P. Cicuta, and M. C. Lagomarsino, Soft Matt. 5, 472 (2009).
[31] A. A. Evans, S. E. Spagnolie, and E. Lauga, Soft Matt. 6, 1737 (2010).
[32] R. Dreyfus, J. Baudry, M. L. Roper, M. Fermigier, H. A. Stone, and J. Bibette, Nature 437, 862 (2005).
[33] W. Gao, S. Sattayasamitsathit, K. M. Manesh, D. Weihs, and J. Wang, J. Am. Chem. Soc. 132, 14403 (2010).
[34] N. Vandenberghe, S. Childress, and J. Zhang, Phys. Fluids 18, 014102 (2006).
[35] S. Spagnolie, L. Moret, M. Shelley, and J. Zhang, Phys. Fluids 22, 041903 (2010).
[36] R. Trouilloud, T. S. Yu, A. E. Hosoi, and E. Lauga, Phys. Rev. Lett. 101, 048102 (2008).
[37] O. S. Pak, T. Normand, and E. Lauga, Phys. Rev. E 81, 036312 (2010).
[38] R. M. Macnab, Proc. Natl. Acad. Sci. USA 74, 221 (1977).
[39] H. C. Berg, Phys. Today 53, 24 (2000).
[40] H. C. Berg, E. coli in Motion (Springer-Verlag, New York, 2004).
[41] A. T. Chwang and T. Y. Wu, Proc. Roy. Soc. Lond. B 178, 327 (1971).
[42] W. Schreiner, J. Biomech. 4, 73 (1971).
[43] J. J. L. Higdon, J. Fluid Mech. 94, 331 (1979).
[44] S. F. Goldstein, N. W. Charon, and J. A. Kreiling, Proc. Natl. Acad. Sci. USA 91, 3433 (1994).
[45] J. W. Shaevitz, J. Y. Lee, and D. A. Fletcher, Cell 122, 941 (2005).
[46] K. E. Summers and I. R. Gibbons, Proc. Natl. Acad. Sci. USA 68, 3092 (1971).
[47] C. J. Brokaw, Science 243, 1593 (1989).
[48] S. Camalet and F. Julicher, New J. Phys. 2, 1 (2000).
[49] J. J. L. Higdon, J. Fluid Mech. 90, 685 (1979).
[50] M. Werner and L. W. Simmons, Biol. Rev. 83, 191 (2008).
[51] J. R. Blake and M. A. Sleigh, Biol. Rev. Camb. Phil. Soc. 49, 85 (1974).
[52] S. Gueron, K. Levit-Gurevich, N. Liron, and J. J. Blum, Proc. Natl. Acad. Sci. USA 94, 6001 (1997).
[53] S. Gueron and K. Levit-Gurevich, Proc. Natl. Acad. Sci. USA 96, 12240 (1999).
[54] B. Guirao and J. F. Joanny, Biophys. J. 92, 1900 (2007).
[55] L. E. Becker, S. A. Koehler, and H. A. Stone, J. Fluid Mech. 490, 15 (2003).
[56] D. Tam and A. E. Hosoi, Phys. Rev. Lett. 98, 068105 (2007).
[57] J. E. Avron and O. Raz, New J. Phys. 10, 063016 (2008).
[58] A. Najafi and R. Golestanian, Phys. Rev. E 69, 062901 (2004).
[59] A. Najafi and R. Golestanian, J. Phys. - Cond. Mat. 17, S1203 (2005).
[60] R. Dreyfus, J. Baudry, and H. A. Stone, Europ. Phys. J. B 47, 161 (2005).
[61] R. Golestanian, Phys. Rev. Lett. 105, 018103 (2010).
[62] J. E. Avron, O. Kenneth, and D. H. Oaknin, New J. Phys. 7, 234 (2005).
[63] A. Najafi and R. Zargar, Phys. Rev. E 81, 067301 (2010).
[64] M. Iima and A. S. Mikhailov, Epl 85, 44001 (2009).
[65] R. M. Thaokar, H. Schiessel, and I. M. Kulic, European Phys. J. B 60, 325 (2007).
[66] A. M. Leshansky and O. Kenneth, Phys. Fluids 20, 063104 (2008).
[67] A. M. Leshansky, O. Kenneth, O. Gat, and J. E. Avron, New J. Phys. 9, 126 (2007).
[68] S. Spagnolie and E. Lauga, Phys. Fluids 22, 081902 (2010).
8
[69] C. H. Wiggins and R. E. Goldstein, Phys. Rev. Lett. 80, 3879 (1998).
[70] C. P. Lowe, Philos. Trans. R. Soc. London, Ser. B 358, 1543 (2003).
[71] M. C. Lagomarsino, F. Capuani, and C. P. Lowe, J. Theor. Biol. 224, 215 (2003).
[72] E. Lauga, Phys. Rev. E 75, 041916 (2007).
[73] T. S. Yu, E. Lauga, and A. E. Hosoi, Phys. Fluids 18, 091701 (2006).
[74] C. W. Wolgemuth, T. R. Powers, and R. E. Goldstein, Phys. Rev. Lett. 84, 1623 (2000).
[75] S. Lim and C. S. Peskin, SIAM J. Sci. Comput. 25, 2066 (2004).
[76] H. Wada and R. R. Netz, Europhys. Lett. 75, 645 (2006).
[77] M. Manghi, X. Schlagberger, and R. R. Netz, Phys. Rev. Lett. 96, 068101 (2006).
[78] N. Coq, O. du Roure, J. Marthelot, D. Bartolo, and M. Fermigier, Phys. Fluids. 20, 051703 (2008).
[79] B. Qian, T. R. Powers, and K. S. Breuer, Phys. Rev. Lett. 1, 078101 (2008).
[80] A. Cebers, Curr. Opin. Colloid Int. Sci. 10, 167 (2005).
[81] M. Roper, R. Dreyfus, J. Baudry, M. Fermigier, J. Bibette, and H. A. Stone, J. Fluid Mech. 554, 167 (2006).
[82] E. Gauger and H. Stark, Phys. Rev. E 74, 021907 (2006).
[83] E. E. Keaveny and M. R. Maxey, J. Fluid Mech. 598, 293 (2008).
[84] M. Roper, R. Dreyfus, J. Baudry, M. Fermigier, J. Bibette, and H. A. Stone, Proc. Roy. Soc. A 464, 877 (2008).
[85] S. Childress and R. Dudley, J. Fluid Mech. 498, 257 (2004).
[86] N. Vandenberghe, J. Zhang, and S. Childress, J. Fluid Mech. 506, 147 (2004).
[87] S. Alben and M. Shelley, Proc. Natl. Acad. Sci. USA 102, 11163 (2005).
[88] X. Y. Lu and Q. Liao, Phys. Fluids 18, 098104 (2006).
[89] S. Childress, in Proceedings of the 2004 International Workshop on Mathematical Fluid Dynamics and Applications, edited
by S.-I. S. J. R. Kweon, S.-C. Kim (2004), pp. 9–21.
[90] E. Lauga, Phys. Fluids 19, 061703 (2007).
[91] M. Roper, Ph.D. thesis, Harvard University, Cambridge, MA (2007).
[92] D. Gonzalez-Rodriguez and E. Lauga, J. Phys. - Cond. Mat. 21, 204103 (2009).
[93] N. H. Mendelson, A. Bourque, K. Wilkening, K. R. Anderson, and J. C. Watkins, J. Bacteriol. 181, 600 (1999).
[94] X. L. Wu and A. Libchaber, Phys. Rev. Lett. 84, 3017 (2000).
[95] C. Dombrowski, L. Cisneros, S. Chatkaew, R. E. Goldstein, and J. O. Kessler, Phys. Rev. Lett. 93, 098103 (2004).
[96] L. H. Cisneros, R. Cortez, C. Dombrowski, R. E. Goldstein, and J. O. Kessler, Exp. Fluids 43, 737 (2007).
[97] A. Sokolov, I. S. Aranson, J. O. Kessler, and R. E. Goldstein, Phys. Rev. Lett. 98, 158102 (2007).
[98] A. Sokolov and I. S. Aranson, Phys. Rev. Lett. 103, 148101 (2010).
[99] S. Rafaı, L. Jibuti, and P. Peyla, Phys. Rev. Lett. 104, 098102 (2010).
[100] E. Lauga and D. Bartolo, Phys. Rev. E 78, 030901 (2008).
[101] G. P. Alexander and J. M. Yeomans, Euro. Phys. Lett. 83, 34006 (2008).
[102] D. Bartolo and E. Lauga, Phys. Rev. E 81, 026312 (2010).
[103] J. B. Shukla, B. R. P. Rao, and R. S. Parihar, J. Biomech. 11, 15 (1978).
[104] D. F. Katz and S. A. Berger, Biorheol. 17, 169 (1980).
[105] S. S. Suarez and X. B. Dai, Biol. Reprod. 46, 686 (1992).
[106] E. Lauga, Phys. Fluids 19, 083104 (2007).
[107] H. C. Fu, T. R. Powers, and H. C. Wolgemuth, Phys. Rev. Lett. 99, 258101 (2007).
[108] H. C. Fu, C. W. Wolgemuth, and T. R. Powers, Phys. Rev. E 78, 041913 (2008).
[109] H. C. Fu, C. W. Wolgemuth, and T. R. Powers, Phys. Fluids 21, 033102 (2009).
[110] E. Lauga, Europhys. Lett. 86, 64001 (2009).
[111] J. Teran, L. Fauci, and M. J. Shelley, Phys. Rev. Lett. 104, 038101 (2010).
[112] R. B. Bird, R. C. Armstrong, and O. Hassager, Dynamics of Polymeric Liquids. Second Edition. Vol. 1: Fluid Mechanics.
(Wiley-Interscience, New York, NY, 1987).
[113] R. B. Bird, C. F. Curtiss, R. C. Armstrong, and O. Hassager, Dynamics of Polymeric Liquids. Second Edition. Vol. 2:
Kinetic Theory. (Wiley-Interscience, New York, NY, 1987).
[114] T. Normand and E. Lauga, Phys. Rev. E 78, 061907 (2008).
|
1604.00412 | 2 | 1604 | 2016-06-05T20:18:22 | Stabilizing membrane domains antagonizes n-alcohol anesthesia | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.CB",
"q-bio.NC"
] | Diverse molecules induce general anesthesia with potency strongly correlated both with their hydrophobicity and their effects on certain ion channels. We recently observed that several n-alcohol anesthetics inhibit heterogeneity in plasma membrane derived vesicles by lowering the critical temperature ($T_c$) for phase separation. Here we exploit conditions that stabilize membrane heterogeneity to further test the correlation between the anesthetic potency of n-alcohols and effects on $T_c$. First we show that hexadecanol acts oppositely to n-alcohol anesthetics on membrane mixing and antagonizes ethanol induced anesthesia in a tadpole behavioral assay. Second, we show that two previously described `intoxication reversers' raise $T_c$ and counter ethanol's effects in vesicles, mimicking the findings of previous electrophysiological and behavioral measurements. Third, we find that hydrostatic pressure, long known to reverse anesthesia, also raises $T_c$ in vesicles with a magnitude that counters the effect of butanol at relevant concentrations and pressures. Taken together, these results demonstrate that $\Delta T_c$ predicts anesthetic potency for n-alcohols better than hydrophobicity in a range of contexts, supporting a mechanistic role for membrane heterogeneity in general anesthesia. | physics.bio-ph | physics | 000 Volume: 00 Month Year 1–20
1
Stabilizing membrane domains antagonizes n-alcohol anesthesia
B.B. Machta, E. Gray, M. Nouri, N.L.C. McCarthy,
E.M. Gray, A.L. Miller, N.J. Brooks and S.L. Veatch
6
1
0
2
n
u
J
5
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
2
1
4
0
0
.
4
0
6
1
:
v
i
X
r
a
© 2013 The Authors
0006-3495/08/09/2624/12
$2.00
2
ABSTRACT
Diverse molecules induce general anesthesia with potency strongly correlated both with their
hydrophobicity and their effects on certain ion channels. We recently observed that several n-alcohol
anesthetics inhibit heterogeneity in plasma membrane derived vesicles by lowering the critical tem-
perature (Tc) for phase separation. Here we exploit conditions that stabilize membrane heterogeneity
to further test the correlation between the anesthetic potency of n-alcohols and effects on Tc. First
we show that hexadecanol acts oppositely to n-alcohol anesthetics on membrane mixing and antag-
onizes ethanol induced anesthesia in a tadpole behavioral assay. Second, we show that two previ-
ously described 'intoxication reversers' raise Tc and counter ethanol's effects in vesicles, mimick-
ing the findings of previous electrophysiological and behavioral measurements. Third, we find that
hydrostatic pressure, long known to reverse anesthesia, also raises Tc in vesicles with a magnitude
that counters the effect of butanol at relevant concentrations and pressures. Taken together, these
results demonstrate that ∆Tc predicts anesthetic potency for n-alcohols better than hydrophobicity in
a range of contexts, supporting a mechanistic role for membrane heterogeneity in general anesthesia.
INTRODUCTION
The potencies of many general anesthetics are roughly proportional to their oil:water partition coeffi-
cient over more than five orders of magnitude in overall concentration (1). This Meyer-Overton correla-
tion suggests membrane involvement, and anesthetics have been shown to decrease lipid chain ordering,
lower the main chain melting temperature, and increase membrane spontaneous curvature, fluidity, and
conductance (2–4). However, these effects are small (5) and often cannot account for those molecules
which deviate from Meyer-Overton (6). Most recent attention focuses on the ion channels known to be
most sensitive to these compounds (7), where extensive structural work (8–10) suggests that anesthetic
effects are mediated by specific residues in hydrophobic, membrane spanning regions. High potency
general anesthetics such as etomidate and barbiturates are more effective at potentiating channels and
producing anesthesia than predicted from their hydrophobicity alone, and evidence is accumulating that
these compounds bind directly and specifically to channels at the interface between subunits (11). Many
researchers also favor a direct binding mechanism of anesthetic action for lower potency anesthetics,
including the n-alcohol anesthetics investigated here (8). However, channels are proposed to contain
more numerous binding sites for these compounds (11–13), each with low affinity, leaving open the
possibility that these anesthetics interact with channels more as a solvent than as a ligand.
000 00(00) 1–20
3
Our understanding of the structure and function of the animal plasma membrane has grown dramat-
ically since most membrane theories of anesthesia were put forward. It is now appreciated that animal
plasma membranes have a thermodynamic tendency to separate into coexisting liquid domains, some-
times referred to as 'lipid rafts' or 'lipid shells' (14, 15) and that this heterogeneity localizes and regulates
ion channels, sometimes in a subtype specific manner (16). Much of this regulation likely arises from
the membrane's unusual thermodynamic properties. Cholesterol containing membranes of purified lipids
can support two distinct liquid phases (17), and giant plasma membrane vesicles (GPMVs) isolated from
mammalian cell lines display analogous phase coexistence at low temperature (18). Remarkably, GPMVs
are near the critical point of this transition (19), a non-generic region of phase space distinguished by
large correlation times and large but finite domain sizes that requires fine tuning both composition and
temperature in synthetic systems (20).
We recently found (21) that incubating several general anesthetics with isolated GPMVs low-
ered their critical temperatures (Tc) in a way that scales well with their anesthetic dose as pre-
viously measured in tadpole loss of righting reflex (LRR) assays (22). While our assay mea-
sures a change in Tc somewhat below growth temperature, we predict
that at higher tempera-
tures these treatments would destabilize sub-micron liquid domains both in vesicles and the intact
plasma membranes from which they were derived. While suggestive, we wanted to rule out the
possibility that our observed correlation is derivative of a more fundamental correlation of both
∆Tc and anesthetic potency with hydrophobicity. As a first step towards this, we demonstrated
that two hydrophobic but non-anesthetic analogs of general anesthetics did not affect Tc at con-
centrations where Meyer-Overton predicts they would (21). Here we explore a more direct chal-
lenge to the connection between changes in Tc and anesthetic potency by investigating the anes-
thetic effects of hydrophobic compounds and conditions that raise transition temperatures in GPMVs.
MATERIALS AND METHODS
Giant plasma membrane vesicle (GPMV) measurements: RBL-2H3 cells (23) were maintained in MEM
media with 20% FBS and 0.1% Gentamycin at 37◦C in 5% CO2. XTC-2 cells (24) were maintained
in L-15 Media diluted 1:1.5 with water for amphibian cells with 10% FBS, sodium bicarbonate (2.47
g/liter), pen strep (100 units/ml) at room temperature in 5% CO2. Freshly seeded cells were incubated in
000 00(00) 1–20
4
complete media for at least 18h at the growth temperature indicated prior to GPMV isolation. All cul-
ture reagents were purchased from Fisher Scientific (Hampton, New Hampshire). Other reagents were
purchased from Sigma Aldrich (St. Louis, MO) at the highest available purity unless otherwise indicated.
GPMVs from RBL cells were prepared through incubation with low concentrations of dithiothreitol
(DTT, 2mM) and formaldehyde (25mM) in the presence of calcium (2mM) for 1h as described previously
(21). GPMVs isolated from RBL cells contain 20-40 mol% cholesterol, sphingolipids, phospholipids,
and gangliosides typically associated with the plasma membrane, along with many transmembrane and
peripheral plasma membrane proteins (18, 25, 26). For XTC-2 derived GPMVs, the vesiculation buffer
was diluted 1:5 while maintaining calcium, formaldehyde, and DTT concentrations, and cells were incu-
bated for at least 2h at room temperature. Prior to GPMV formation, cells were labeled with DiI-C12
(Life Technologies, Carlsbad, CA; 2µg/ml in 1% methanol) for 10min at room temperature. GPMVs
probed at atmospheric pressure were imaged on an inverted microscope (IX81; Olympus, Center Valley,
PA) with a 40x air objective (0.95 NA), epi-illumination using an Hg lamp and Cy3 filter set (Chroma
Technology, Bellows Falls, VT). Temperature was controlled using a home built Peltier stage described
previously (21) coupled to a PID controller (Oven Industries, Mechanicsburg, PA), and images were
recorded using a sCMOS camera (Neo; Andor, South Windsor, CT).
GPMV suspensions with hexadecanol were prepared either using super-saturated solutions or equi-
librated solutions. To make super-saturated solutions, hexadecanol was suspended in either DMSO
or ethanol using volumes corresponding to the final concentration indicated in figures, then mixed
directly with the GPMV suspension in aqueous buffer while mixing. The maximum DMSO con-
centration used was 0.5% v/v, and previous work demonstrates that Tc is not affected by DMSO
treatment alone (21). Equilibrated solutions were prepared by first adding a concentrated hexade-
canol stock directly to the GPMV suspension such that it precipitated out of solution. Then, the
desired volume of ethanol was added and the solution mixed to facilitate the resuspension of hex-
adecanol. In all cases where ethanol and hexadecanol were added in combination, they were main-
tained at a 120mM ethanol to 30µM hexadecanol ratio, as indicated in figures. Equilibrated solutions
could be made for the following: 60mM:1.5µM, 120mM:3µM, 180mM:4.5µM, and 240mM:6µM
(ethanol:hexadecanol), although for higher concentrations some small fraction of hexadecanol lacked
solubility. Only super-saturated solutions could be made for 600mM ethanol:15µM hexadecanol.
000 00(00) 1–20
5
Fig. 1. Determination of the average critical temperature or pressure of DiIC12 labeled GPMVs through fluorescence
imaging. (A) Fields containing multiple GPMVs were imaged over a range of temperatures and at fixed pressure, with rep-
resentative subsets of images shown on the left. At high temperature, most GPMVs appear uniform, while an increasing
fraction of vesicles appear phase separated as temperature is lowered, with phase separated vesicles indicated by yellow
arrows. We manually tabulate the fraction of GPMVs that contain two coexisting liquid phases as a function of temperature
from these images, constructing the plot on the right. These points are fit to the sigmoid function described in Methods
to determine the extrapolated temperature where 50% of vesicles contain coexisting liquid phases. (B) Fields contain-
ing multiple GPMVs were imaged over a range of pressures at fixed temperature, and representative subsets of images
are shown on the left. At low pressure, most GPMVs appear uniform, while an increasing fraction of vesicles appear
phase separated as pressure is increased. As with the fixed pressure data in A, these points are fit to the sigmoid func-
tion described in Methods to determine the extrapolated pressure where 50% of vesicles contain coexisting liquid phases.
000 00(00) 1–20
6
Measurements conducted at elevated hydrostatic pressures were made using a custom built
microscopy compatible pressure cell mounted on a Nikon Eclipse TE2000-E inverted microscope as
described previously (27, 28) with a 20x extra long working distance air objective (0.4 NA) and G2-A
filter set (Nikon Instruments, Richmond, UK) The pressure cell temperature was controlled via a circu-
lating water bath. Images were acquired using a sCMOS based camera (Zyla; Andor, Belfast, UK) and
recorded using custom built software with temperature and pressure logging.
GPMV transition temperatures at constant pressure were measured as described previously (21) and
as illustrated in Fig 1A. Briefly, images were acquired of fields of GPMVs over a range of temperatures
such that at least 100 vesicles were detected at each temperature. After imaging, individual vesicles were
identified as having a single liquid phase or two coexisting liquid phases. This information was compiled
into a plot showing the percentage of vesicles with two liquid phases as a function of temperature, which
was fit to a sigmoid function to extrapolate the temperature where 50% of vesicles contained two coex-
, where B is a parameter describing
the width of the transition. The width of the transition within a population of vesicles (∼ 10◦C) is much
broader than the width of the transition for a single GPMV (< 2◦C) (19), most likely due to heterogeniety
in composition between GPMVs (29).
isting liquid phases, % Phase Separated = 100×(cid:16)
1+e−(T−Tc)/B
1 −
(cid:17)
1
We have previously demonstrated that these GPMVs pass through a critical temperature at the tran-
sition, even in the presence of anesthetics, therefore we refer to this temperature as the average critical
temperature (Tc) of the sample. Errors in single measurements of Tc (σTc) are 68% confidence interval
estimates of this parameter determined directly from the fit. For the example shown in Fig 1A, the critical
temperature is Tc=19.5±0.6◦C. We generally observe that average critical temperatures measured in this
way vary between 12◦C < Tc < 27◦C for untreated RBL-derived GPMVs prepared as described above,
depending on the growth conditions of the cells from which the GPMVs were derived (29). XTC-2 cells
produced GPMVs with average critical temperatures of 14.8±0.4◦ and 20.7±0.4◦C. Error bounds for
C where σC is the error in measuring Tc of
the untreated control and σM is the error in measuring Tc of the treated sample. The average critical
pressure in the sample (Pc) at constant temperature was determined by a similar procedure, as illus-
.
As with previous studies varying only temperature (19), we find that the fraction of phase separated
vesicles present in a population of vesicles at a given temperature/pressure does not depend on the
order in which temperature or pressures are sampled, suggesting the transition is fully reversible.
a transition temperature shift (∆Tc) are given by(cid:112)σ2
trated in Fig1B, although fit to a slightly different equation, % Phase Separated = 100×(cid:16)
1+e−(P−Pc)/B
M + σ2
(cid:17)
1
000 00(00) 1–20
7
Tadpole loss of righting reflex (LRR) measurements: Studies with Xenopus laevis tadpoles were con-
ducted in compliance with the US Department of Health and Human Services Guide for the Care and
Use of Laboratory Animals and were approved by the University of Michigan IACUC. Xenopus laevis
embryos were collected, fertilized, and dejellied as described previously (30). Embryos were stored at
room temperature in 0.1X MMR (1X MMR = 100mM NaCl, 2mM KCl, 2mM CaCl2, 1mM MgCl2,
5mM Hepes, pH 7.4) and allowed to develop to swimming tadpole stage (stage 43-45).
The LRR response was determined by placing tadpoles in a clean glass container filled with 50ml
well water and the specified concentration of n-alcohols. At 10min intervals, tadpoles were probed with
a smooth glass rod and their responses recorded. Measurements with ethanol or ethanol and hexade-
canol in equilibrated solutions were conducted on three separate occasions each with five tadpoles per
condition, totaling fifteen tadpoles per condition. Some measurements were conducted double-blind in
order to avoid possible systematic bias in the behavioral scoring. Fewer tadpoles were probed for the
other n-alcohol combinations investigated. Error bars on LRR measurements are 68% confidence inter-
vals with binomial errors calculated according to: LB = 1 − BetaInv(0.32/2, n − nk, nk + 1) and
U B = 1 − BetaInv(1 − 0.32/2, n − nk + 1, nk) where LB and U B are the lower and upper bounds of
the confidence interval, n is the number of tadpoles investigated, k is the fractional LRR, and BetaInv is
the Beta inverse cumulative distribution function.
1
To estimate AC50, LRR measurements at a range of ethanol concentrations were fit to the form
LRR = 1 −
1+e−([EtOH]−AC50)/B . Errors reported are 68% confidence intervals of parametric uncer-
tainty in AC50. These are calculated by inverting the expectation value of the Fisher Informa-
tion and then taking the square root of its diagonal entry. Error bars are much larger for hex-
adecanol containing titrations because all data is taken below the extrapolated AC50 concentration.
RESULTS
Fig. 2A shows that incubating super-saturated solutions of hexadecanol (n=16 alcohol) with isolated
Rat Basophilic Leukemia (RBL) and XtC-2 derived GPMVs acts to raise Tc. Here, GPMVs derived
from both mammalian (RBL) and Xenopus laevis (XtC-2) cells were used to demonstrate robustness
of this effect across species, since later measurements are conducted in Xenopus laevis tadpoles. The
sign change in ∆Tc for hexadecanol compared to shorter chain n-alcohols mirrors past work which
found that hexadecanol acted to increase chain order within native isolated membranes, while shorter
000 00(00) 1–20
8
anesthetic n-alcohols acted to decrease chain order. (31). We find that the effect on Tc is approximately
additive when ethanol (n=2 alcohol) and hexadecanol are used in combination, with 3µM hexadecanol
required to counter the effect of an AC50 of ethanol (120mM) over a range of ethanol and hexadecanol
concentrations (Fig. 2A,B).
Fig. 2. Hexadecanol raises Tc in GPMVs from rat (RBL) and Xenopus (XtC-2) cell lines and can counteract the Tc low-
ering effects of ethanol. (A) Values indicate the average shift in Tc (∆Tc) in a population of vesicles upon treatment with
the compounds indicated. Solutions containing hexadecanol were prepared to be super-saturated as described in Methods.
Each point represents a single measurement and error bounds represent the 68% confidence interval on the extrapolated
∆Tc. (B) Plots showing fraction of phase separated vesicles vs. temperature for the three points inside the gray box in A.
000 00(00) 1–20
9
Fig. 3. (A) (Upper panel) Tadpole loss of righting reflex (LRR) for a titration of ethanol alone and combinations of ethanol
and hexadecanol (EtOH+Hex) or ethanol and tetradecanol (EtOH+Tet) measured after 1h incubation in equilibrated solu-
tions. At a given ethanol concentration,the presence of hexadecanol increases the fraction of tadpoles which respond to
stimulus. (Lower panel) ∆Tc in RBL derived GPMVs for identical titrations of ethanol and EtOH+Hex. All solutions con-
tain the ethanol concentration indicated by the lower horizontal axis. Red circle points additionally contain hexadecanol
concentrations indicated by the upper horizontal axis, and green triangle points additionally contain either 5 or 10µM tetrade-
canol. (B) Time-course of LRR for one ethanol and ethanol+hexadecanol combination. (C) (Left) Points in A replotted as
LRR vs ∆Tc, including additional experiments with other n-alcohol combinations as indicated in the legend. (Center) LRR
plotted vs. aggregate hydrophobicity, tabulated by summing the concentration of each n-alcohol present normalized by its
AC50 (22), using 3µM as a proxy AC50 for hexadecanol and 5µM as a proxy AC50 for tetradecanol. (Right) LRR plot-
ted vs. the net anesthetic concentration, tabulated by summing the concentration of each n<14 alcohol anesthetic present
normalized by its AC50 (22). (Top panels) In each case the dashed line represents a best fit to the data with R2 values
giving the fraction of explained variance and f values comparing the quality of the fit to a null model where all points
have the same probability. All fits are linear and constrained to go through the origin. (Bottom Panels) In each case the
black line is fit to all conditions that exclude hexadecanol, the red line is fit to all conditions that include hexadecanol,
and the gray dashed line is fit to all points as in the upper panels. These fits are substantially different from each other
except when plotted vs ∆Tc, indicating that ∆Tc has the most quantitative predictive power across chemical species.
Mimicking results from our previous study with n-alcohols (21), we again find that relatively small
concentrations of both hexadecanol and ethanol lead to a steep change in Tc, while larger concentrations
lead to a less steep and approximately linear regime. This nonlinearity implies an interaction between
000 00(00) 1–20
10
the added molecules, but we cannot specify the nature of their interaction from our measurements. It
could be that the concentration of n-alcohol in the membrane becomes a sub-linear function of their con-
centration in the bulk. Past work has demonstrated that n-alcohol membrane partition coefficients can
depend on n-alcohol concentration (32), membrane composition (32, 33), and temperature (33), and it is
possible that these or related effects could lead to the observed nonlinearity. Another possibility is that
the addition of small molecules could increase the area of the membrane, thereby reducing its tension.
This could have a direct effect on Tc (34), or could alter the membrane's affinity for additional small
molecules. It is also possible that membrane concentration is linear with that of the bulk, but that Tc is a
nonlinear function of n-alcohol concentration in the membrane. We note that ethanol can induce a novel
interdigitated membrane phase in purified phosphatidylcholine membranes when at high enough con-
centration (35, 36) so it is plausible that specific interactions between certain lipids and alcohols could
drive at least some of the nonlinearity that we observe. At AC50, small molecules make up several mol%
of the membrane (22), a relatively high concentration which makes it plausible that n-alcohols interact
directly, or indirectly through membrane mediated mechanisms that can be very sensitive near the critical
point (37).
Our observation that hexadecanol can counteract the effects of ethanol in GPMVs led us to speculate
that hexadecanol could antagonize ethanol anesthesia. To test this hypothesis, we examined the tadpole
loss of righting reflex (LRR) in Xenopus laevis tadpoles (Fig.3A). We also measured the extent to which
identical n-alcohol treatments alter Tc in RBL derived GPMVs. Here equilibrated solutions of hexade-
canol and ethanol were used, leading to small differences from the results presented in Fig 2. In ethanol
alone, 50% of tadpoles did not exhibit a righting reflex upon inverting with a glass rod at 118±15mM
ethanol in agreement with previous reports of 120mM (22). Tadpoles incubated in ethanol and hexade-
canol remained alert to much higher ethanol concentrations, more than doubling the extrapolated AC50
value to 294±61mM. Co-incubation with tetradecanol (n=14) had little effect on either GPMV transition
temperatures or on the tadpole LRR, suggesting that competitive inhibition at a putative ethanol binding
site is not the mechanism through which hexadecanol reverses the effects of ethanol on LRR.
Fig.3B shows the time dependent effects of ethanol and ethanol-hexadecanol mixtures on tadpole
LRR. LRR is not initially affected by the presence of hexadecanol, but instead reduces LRR over the
span of 1h compared to those incubated in ethanol alone. Experiments were not extended beyond 1h
as we observed some adaptation of ethanol treated tadpoles beyond this time-frame. The slow onset
of hexadecanol action is consistent with past work using radio-labeled n-alcohols, which demonstrated
000 00(00) 1–20
11
that their absorption dynamics depend on carbon length, with longer n-alcohols requiring longer times
for incorporation (31). We note that the apparent difference in LRR for ethanol+hexadecanol vs. ethanol
alone is not significant at early time-points, nor is the apparent upward trend in LRR vs. time for tadpoles
treated with ethanol alone. In Fig.3C we plot LRR vs ∆Tc, aggregate hydrophobicity, or anesthetic con-
centration for a range of conditions. We find that ∆Tc captures substantially more of the variance in LRR
across trials (R2 = .82) than summed anesthetic potency (R2 = .53), while aggregate hydrophobicity
has no predictive power (R < 0). This is partially due to our carefully chosen treatments, many of which
are hydrophobic but raise critical temperatures and antagonize LRR. Furthermore, we find that ∆Tc pre-
dicts LRR with the same function whether for treatments containing hexadecanol or treatments without
hexadecanol, while both aggregate hydrophobicity and summed anesthetic potency do not. Together this
demonstrates that ∆Tc predicts LRR across molecular species.
This indicates that ∆Tc is significantly more predictive of anesthetic potency than aggregate
hydrophobicity or anesthetic concentration for the n-alcohol mixtures examined.
The extremely low solubility of hexadecanol presents some experimental difficulties. Ethanol is
required as a co-solvent in these measurements and this prevented a more systematic exploration of
how hexadecanol modulates anesthesia mediated by other n-alcohols and anesthetics. Our reliance
on ethanol as a general anesthetic introduces technical challenges since it has low potency, can
harbor impurities that modulate LRR measurements (31), and produces alternate functional out-
comes when used at high concentrations (38). While a more water soluble Tc raising compound
would alleviate some of these problems, most membrane soluble molecules with reasonable water
solubility either lower GPMV transition temperatures or leave them unchanged. Notable excep-
tions are some detergents (39), which are not suitable for this investigation because they also
permeabilize membranes. Molecules which raise Tc must partition more strongly than the aver-
age component into one low temperature membrane phase (40) and our observations suggest that
this often requires a large hydrophobic interaction area, with consequent low solubility in water.
000 00(00) 1–20
12
Fig. 4. Ro15-4513 and DHM block the acute toxicity and intoxicating effect of ethanol. Each raise Tc and can-
cel
the effects of ethanol when added to GPMVs at
the same concentration at which they are effective in vivo.
Several other compounds are reported to reverse the intoxicating effects of ethanol in both cultured
neurons and intact organisms, and their effects on RBL derived GPMVs are shown in Fig 4. 100nM
RO15-4513, a therapeutic agent used to reverse acute ethanol toxicity, has been demonstrated to reverse
effects of 30mM ethanol in the GABA current in oocytes (41). We find that 100nM RO15-4513 raises
critical temperatures in GPMVs by 1.2 ± 0.4◦C. Similarly, 3µM Dihydromyricetin (DHM) was recently
reported to antagonize the effects of 60mM ethanol on the GABA-induced current in acutely dissociated
rat hippocampal slices as well as on behavioral assays in vivo (42). We find that 3µM DHM raises criti-
cal temperatures by 1.3 ± 0.5◦C. In each of these cases, the effect of these compounds on Tc appears to
saturate at concentrations well below their solubility. While we have no mechanistic explanation for this
saturation, we note that it is mirrored in past behavioral measurements (41, 42) which find that they are
only able to counter ethanol effects to concentrations well below its AC50 of 120mM.
While the correlation between ∆Tc and anesthetic function is robust for the compounds investi-
gated, we note that an additional compound, menthol, raises Tc by nearly 2◦C when added to RBL
derived GPMVs at 100µM (43), yet acts as an anesthetic in tadpole LRR measurements at some-
what lower concentration (AC50 ≈20µM) (44). This could be due to menthol having specific inter-
actions with GABAA receptors, as has been proposed previously (44). It is also possible that men-
thol has different effects on Xenopus laevis neuronal vs. mammalian immune plasma membranes.
000 00(00) 1–20
13
Fig. 5. (A) The fraction of vesicles which are macroscopically phase separated is plotted as a function of hydrostatic
pressure at three different temperatures, both for control vesicles and vesicles incubated with 12mM bBtOH. In each
case, increasing the pressure leads to an increase in the fraction of vesicles which are macroscopically phase sepa-
rated. (B) Tc is raised with increasing hydrostatic pressure in both control GPMVs and GPMVs incubated in butanol
(BtOH). Here, 240±30 bar of hydrostatic pressure is required to reverse the effects of 12mM BtOH (shaded region).
Closed symbols are obtained by extrapolating to find Tc from data acquired at constant pressure while open symbols
are obtained by extrapolating to find Pc at constant temperature. At temperatures above Tc most vesicles are com-
posed of a macroscopically uniform single liquid, while below Tc most are separated into two co-existing liquid phases.
DISCUSSION
We also investigated the effects of hydrostatic pressure on GPMVs in the presence and absence of
butanol (n=4 alcohol). It has long been known that 150-200 bar of pressure reverses animal anesthe-
sia (45). Here we show that the fraction of vesicles with two phase coexistence increases monotonically
000 00(00) 1–20
14
with pressure, for vesicles incubated with and without butanol (Fig 5A). We observe that 240±30 bar is
needed to counter the effects of one AC50 of butanol (12mM (22)) on the Tc of RBL derived GPMVs
(Fig 5B). Pressure reversal of general anesthesia is not accounted for by current models involving direct
binding of anesthetics to channels because protein conformational equilibria are typically sensitive only
to much larger pressures (>1000 bar) (46, 47). These smaller pressures can have large effects on some
membrane properties, and past work has shown that 200 bar raises the main chain melting temperature
in synthetic single component lipid systems by ∼ 4.5◦C (48, 49), even more than is observed here for
the miscibility transition in GPMVs (2-3◦C).
While our results are not sufficient to specify a mechanism, they do strongly suggest that n-alcohol
anesthetic inhibition of membrane heterogeneity plays an important role in mediating anesthesia, likely
by interfering with normal membrane regulation of ion channels. We have previously argued that the
presence of a critical point in GPMVs near or below room temperature leads to extended and dynamic
composition fluctuations at growth temperature (typically 37◦C), both in isolated vesicles and in the
intact plasma membranes from which they are derived (19, 50). Theory and experiment suggest that
the size of fluctuations varies as (T − Tc)−1 (19, 51), so that lowering Tc through the addition of an
n-alcohol anesthetic is expected to reduce the size of compositional heterogeneity at fixed growth temper-
ature (Fig. 6). This suggests several plausible mechanisms through which changes in plasma membrane
Tc could modulate channel function. First, reducing the size of membrane structures could in princi-
ple increase (or decrease) accessibility of native regulators of channel function such as neurosteroids,
phosphoinositides and G-proteins (16). Second, different membrane domains are expected to have dif-
fering physical properties such as hydrophobic thickness, lateral pressure profiles, viscosity, and bending
rigidity (52–56). Destabilizing these membrane domains could act to modulate the response of channels
directly by stabilizing different internal protein states. Thirdly, membrane domains are argued to play
important roles in organizing channels at neuronal synapses (16) and destabilizing these domains could
lead to altered localization of channels with disruption of higher level neuronal functions. Each of these
mechanistic possibilities have distinct experimental signatures that can be explored in future studies.
000 00(00) 1–20
15
Fig. 6. Anesthetics lower the transition temperature of a biologically tuned critical point which could lead to mis-regulation
of ion channels and other membrane bound proteins. (A) Schematic phase diagrams for the plasma membrane of an untreated
and n-alcohol treated cell. Guided by experiments (19) we hypothesize that the plasma membrane lies in the white region
where lipids, proteins and other membrane components are well mixed macroscopically into a single two dimensional liquid
phase. However, due to their close proximity to the critical point (star), thermal fluctuations lead to relatively large domains
enriched in particular components. When cooled into the gray two phase region, GPMVs separate into two coexisting liquid
phases termed liquid-ordered and liquid-disordered. Several n-alcohol general anesthetics lower the critical temperature of
the membrane (21), changing the distance above the critical point, T − Tc. Here we also show that treatments which antag-
onize anesthetic action raise critical temperatures reversing the effects of n-alcohols on Tc. (B) Under normal conditions
a hypothetical ion channel (large blue inclusion) has a tendency to inhabit relatively large domains enriched in particular
lipids and proteins. When the membrane is taken away from the critical point, the structure of these domains is altered,
possibly leading to changes in ion channel gating and function through a variety of mechanisms discussed in the text.
One naive prediction of our model is not borne out by experiment, namely that the functional effects
of anesthetics are not readily reversed by lowering ambient temperature (57), even though this should
reverse anesthetic effects on the parameter T − Tc. While this finding is at odds with a simple interpreta-
tion of our results, we note that other biological functions have been shown to be explicitly temperature
compensated (58, 59). In these examples individual reaction rates often display sharp temperature sensi-
tivity, but the effects of these changes on certain important axes (e.g. the timing of the circadian clock)
roughly cancel, leaving function robust (60). Thus, while many cellular processes relevant to neuronal
000 00(00) 1–20
16
function have sensitive temperature dependence (61–63), we expect that organisms, especially those able
to live over a range of temperatures, have evolved to a regime where neural function is temperature com-
pensated (63) and is thus surprisingly insensitive to changes in temperature. However, we might still
expect that triggering only temperature's effects on membrane criticality in an uncompensated way, as
we argue occurs with n-alcohol anesthesia, might lead to dramatic changes in function (60). We note that
past studies which demonstrate a lack of temperature dependence on anesthetic potency have conducted
behavioral measurements in temperature adapted animals (64, 65), where other studies have found evi-
dence for concurrent changes in membrane composition (66, 67). Acute temperature changes mimicking
or exceeding those produced by relevant concentrations of anesthetic (±4◦C) do produce behavioral
changes in animals ranging from lethargy to death (64).
Finally, while the evidence for anesthetic-channel interactions is significant, several features may
suggest the interaction is more akin to two dimensional solvent-solute rather than binding site-ligand
interactions. Some channels contain multiple proposed binding sites for low potency anesthetics, each
with low affinity but diffusion limited association rates (12), with clinically relevant concentrations of
anesthetic as high as several mol% of the membrane (68). Additionally, many anesthetic sensitive chan-
nels are also sensitive to membrane properties, and in particular cholesterol modulation both in recon-
stituted (69) and cellular systems (70), suggesting a commonality with broader regulation by membrane
domains. Our results suggest that these effects on channel function are likely to be altered by anes-
thetics through their effects on membrane mixing even without binding of anesthetic to protein targets.
CONCLUSION
Overall, the results presented here demonstrate that a condition's effect on GPMV Tc is more pre-
dictive of anesthetic potency than hydrophobicity. By exploiting rare conditions where ∆Tc > 0 we
have shown that behavioral measures and existing electrophysiological assays for anesthesia are remark-
ably tied to the membrane's thermodynamic propensity to form small domains. While our results do
not suggest a specific mechanism through which these conditions reverse anesthesia, they do support the
hypothesis that at least some n-alcohol targets may be influenced through effects on membrane criticality.
Acknowledgements We thank Keith Miller and James Sethna for helpful converstaions, Ned Wingreen
for a close reading of the manuscript and Margaret Burns, Jing Wu, and Kathleen Wisser for assis-
tance with some experiments. Research was funded by the NIH (R01 GM110052 to SLV and R01
000 00(00) 1–20
GM112794 to ALM), startup funds from the University of Michigan (to SLV), EPSRC Programme grant
(EP/J017566/1 to NJB), a EPSRC Centre for Doctoral Training Studentship awarded by the Institute
of Chemical Biology to NLCM (EP/F500076/1) a Lewis-Sigler Fellowship (BBM) and by NSF PHY
0957573 (BBM).
17
References
1. Meyer, H., 1899. Zur Theorie der Alkoholnarkose. Archiv fr experimentelle Pathologie und Pharmakologie 42:109–118.
2. Gruner, S. M., and E. Shyamsunder, 1991. Is the mechanism of general anesthesia related to lipid membrane spontaneous curvature?
Ann N Y Acad Sci 625:685–97.
3. Wodzinska, K., A. Blicher, and T. Heimburg, 2009. The thermodynamics of lipid ion channel formation in the absence and presence
of anesthetics. BLM experiments and simulations. Soft Matter 5:3319–3330.
4. Ing´olfsson, H. I., and O. S. Andersen, 2011. Alcohol's Effects on Lipid Bilayer Properties. Biophysical Journal 101:847–855.
5. Herold, K. F., R. L. Sanford, W. Lee, M. F. Schultz, H. I. Inglfsson, O. S. Andersen, and H. C. Hemmings, 2014. Volatile anesthetics
inhibit sodium channels without altering bulk lipid bilayer properties. The Journal of General Physiology 144:545–560.
6. Franks, N. P., and W. R. Lieb, 1986. Partitioning of long-chain alcohols into lipid bilayers: implications for mechanisms of general
anesthesia. Proc Natl Acad Sci U S A 83:5116–20.
7. Franks, N. P., and W. R. Lieb, 1994. Molecular and cellular mechanisms of general anaesthesia. Nature 367:607–14.
8. Mihic, S. J., Q. Ye, M. J. Wick, V. V. Koltchine, M. A. Krasowski, S. E. Finn, M. P. Mascia, C. F. Valenzuela, K. K. Hanson, E. P.
Greenblatt, R. A. Harris, and N. L. Harrison, 1997. Sites of alcohol and volatile anaesthetic action on GABA(A) and glycine receptors.
Nature 389:385–389.
9. Borghese, C. M., D. F. Werner, N. Topf, N. V. Baron, L. A. Henderson, S. L. Boehm, Y. A. Blednov, A. Saad, S. Dai, R. A. Pearce,
R. A. Harris, G. E. Homanics, and N. L. Harrison, 2006. An Isoflurane- and Alcohol-Insensitive Mutant GABAA Receptor ?1 Subunit
with Near-Normal Apparent Affinity for GABA: Characterization in Heterologous Systems and Production of Knockin Mice. Journal
of Pharmacology and Experimental Therapeutics 319:208–218.
10. Nury, H., C. Van Renterghem, Y. Weng, A. Tran, M. Baaden, V. Dufresne, J. P. Changeux, J. M. Sonner, M. Delarue, and P. J.
Corringer, 2011. X-ray structures of general anaesthetics bound to a pentameric ligand-gated ion channel. Nature 469:428–+.
11. Forman, S. A., and K. W. Miller, 2016. Mapping General Anesthetic Sites in Heteromeric γ-Aminobutyric Acid Type A Receptors
Reveals a Potential For Targeting Receptor Subtypes. Anesth Analg. in press.
12. Xu, Y., T. Seto, P. Tang, and L. Firestone, 2000. NMR study of volatile anesthetic binding to nicotinic acetylcholine receptors. Biophys
J 78:746–51.
13. Brannigan, G., D. N. LeBard, J. Hnin, R. G. Eckenhoff, and M. L. Klein, 2010. Multiple binding sites for the general anesthetic
isoflurane identified in the nicotinic acetylcholine receptor transmembrane domain. Proceedings of the National Academy of Sciences
107:14122–14127.
14. Simons, K., and E. Ikonen, 1997. Functional rafts in cell membranes. Nature 387:569–72.
15. Anderson, R. G., and K. Jacobson, 2002. A role for lipid shells in targeting proteins to caveolae, rafts, and other lipid domains. Science
296:1821–5.
16. Allen, J. A., R. A. Halverson-Tamboli, and M. M. Rasenick, 2007. Lipid raft microdomains and neurotransmitter signalling. Nat Rev
Neurosci 8:128–140.
000 00(00) 1–20
18
17. Veatch, S. L., and S. L. Keller, 2005. Seeing spots: Complex phase behavior in simple membranes. Biochimica et Biophysica Acta
(BBA) - Molecular Cell Research 1746:172 – 185.
18. Baumgart, T., A. T. Hammond, P. Sengupta, S. T. Hess, D. A. Holowka, B. A. Baird, and W. W. Webb, 2007. Large-scale fluid/fluid
phase separation of proteins and lipids in giant plasma membrane vesicles. Proc Natl Acad Sci U S A 104:3165–70.
19. Veatch, S. L., P. Cicuta, P. Sengupta, A. Honerkamp-Smith, D. Holowka, and B. Baird, 2008. Critical fluctuations in plasma membrane
vesicles. Acs Chemical Biology 3:287–293.
20. Honerkamp-Smith, A. R., P. Cicuta, M. D. Collins, S. L. Veatch, . den Niljs, M. Schick, and S. L. Keller, 2008. Line Tensions,
Correlation Lengths, and Critical Exponents in Lipid Membranes Near Critical Points. Biophysical Journal 95:236–246.
21. Gray, E., J. Karslake, B. B. Machta, and S. L. Veatch, 2013. Liquid general anesthetics lower critical temperatures in plasma membrane
vesicles. Biophys J 105:2751–9.
22. Pringle, M. J., K. B. Brown, and K. W. Miller, 1981. Can the lipid theories of anesthesia account for the cutoff in anesthetic potency
in homologous series of alcohols? Mol Pharmacol 19:49–55.
23. Barsumian, E. L., C. Isersky, M. G. Petrino, and R. P. Siraganian, 1981. IgE-induced histamine release from rat basophilic leukemia
cell lines: isolation of releasing and nonreleasing clones. European Journal of Immunology 11:317–323.
24. Pudney, M., M. G. Varma, and C. J. Leake, 1973. Establishment of a cell line (XTC-2) from the South African clawed toad, Xenopus
laevis. Experientia 29:466–7.
25. Fridriksson, E. K., P. A. Shipkova, E. D. Sheets, D. Holowka, B. Baird, and F. W. McLafferty, 1999. Quantitative Analysis of Phospho-
lipids in Functionally Important Membrane Domains from RBL-2H3 Mast Cells Using Tandem High-Resolution Mass Spectrometry.
Biochemistry 38:8056–8063.
26. Levental, K. R., J. H. Lorent, X. Lin, A. D. S. ad Michal A. Surma, E. A. Stockenbojer, A. A. Gorfe, and I. Levental, 2016.
Polyunsaturated Lipids Regulate Membrane Domain Stability by Tuning Membrane Order. Biophysical Journal 110:1800–1810.
27. McCarthy, N. L. C., O. Ces, R. V. Law, J. M. Seddon, and N. J. Brooks, 2015. Separation of liquid domains in model membranes
induced with high hydrostatic pressure. Chem. Commun. 51:8675–8678.
28. Purushothaman, S., P. Cicuta, O. Ces, and N. J. Brooks, 2015.
Influence of High Pressure on the Bending Rigidity of Model
Membranes. The Journal of Physical Chemistry B 119:9805–9810. 26146795.
29. Gray, E. M., G. Diaz-Vazquez, and S. L. Veatch, 2015. Growth Conditions and Cell Cycle Phase Modulate Phase Transition
Temperatures in RBL-2H3 Derived Plasma Membrane Vesicles. PLoS ONE 10:1–16.
30. Miller, A. L., and W. M. Bement, 2009. Regulation of cytokinesis by Rho GTPase flux. Nat Cell Biol 11:71–77.
31. Miller, K. W., L. L. Firestone, J. K. Alifimoff, and P. Streicher, 1989. Nonanesthetic alcohols dissolve in synaptic membranes without
perturbing their lipids. Proc Natl Acad Sci U S A 86:1084–7.
32. Fraser, D. M., L. C. Van Gorkom, and A. Watts, 1991. Partitioning behaviour of 1-hexanol into lipid membranes as studied by
deuterium NMR spectroscopy. Biochimica et Biophysica Acta (BBA) - Biomembranes 1069:53–60.
33. Trandum, C., P. Westh, K. Jrgensen, and O. G. Mouritsen, 2000. A Thermodynamic Study of the Effects of Cholesterol on the
Interaction between Liposomes and Ethanol. Biophysical Journal 78:2486 – 2492.
34. Portet, T., S. E. Gordon, and S. L. Keller, 2012. Increasing Membrane Tension Decreases Miscibility Temperatures; an Experimental
Demonstration via Micropipette Aspiration. Biophysical Journal 103:L35 – L37.
35. Rowe, E. S., and C. T. A., 1990. Differential Scanning Calorimetric Studies of Ethanol Interactions with Distearoylphosphatidyl-
choline: Transition to the Interdigitated Phase. Biochemistry 29:10398–10404.
36. McIntosh, T. J., H. Lin, S. Li, and C. hsien Huang, 2001. The effect of ethanol on the phase transition temperature and the phase
structure of monounsaturated phosphatidylcholines. Biochimica et Biophysica Acta (BBA) - Biomembranes 1510:219 – 230.
000 00(00) 1–20
19
37. Widom, B., 1967. Plait points in 2 and 3-Component liquid mixtures. J. Chem. Phys. 46:3324.
38. Downes, H., and P. M. Courogen, 1996. Contrasting effects of anesthetics in tadpole bioassays.
Journal of Pharmacology and
Experimental Therapeutics 278:284–296. http://jpet.aspetjournals.org/content/278/1/284.abstract.
39. Zhou, Y., K. N. Maxwell, E. Sezgin, M. Lu, H. Liang, J. F. Hancock, E. J. Dial, L. M. Lichtenberger, and I. Levental, 2013. Bile Acids
Modulate Signaling by Functional Perturbation of Plasma Membrane Domains. Journal of Biological Chemistry 288:35660–35670.
40. Meerschaert, R. L., and C. V. Kelly, 2015. Trace membrane additives affect lipid phases with distinct mechanisms: a modified Ising
model. Eur Biophys J 44:227–33.
41. Wallner, M., H. J. Hanchar, and R. W. Olsen, 2006. Low-dose alcohol actions on α4β3γ GABAA receptors are reversed by the
behavioral alcohol antagonist Ro15-4513. Proceedings of the National Academy of Sciences 103:8540–8545.
42. Shen, Y., A. K. Lindemeyer, C. Gonzalez, X. M. Shao, I. Spigelman, R. W. Olsen, and J. Liang, 2012. Dihydromyricetin As a Novel
Anti-Alcohol Intoxication Medication. The Journal of Neuroscience 32:390–401.
43. Raghunathan, K., A. Ahsan, D. Ray, M. K. Nyati, and S. L. Veatch, 2015. Membrane Transition Temperature Determines Cisplatin
Response. PLoS ONE 10:e0140925.
44. Watt, E. E., B. A. Betts, F. O. Kotey, D. J. Humbert, T. N. Griffith, E. W. Kelly, K. C. Veneskey, N. Gill, K. C. Rowan, A. Jenkins, and
A. C. Hall, 2008. Menthol shares general anesthetic activity and sites of action on the GABAA receptor with the intravenous agent,
propofol. European Journal of Pharmacology 590:120 – 126.
45. Johnson, F. H., and E. A. Flagler, 1950. Hydrostatic pressure reversal of narcosis in tadpoles. Science 112:91–2.
46. Moss, G. W., W. R. Lieb, and N. P. Franks, 1991. Anesthetic inhibition of firefly luciferase, a protein model for general anesthesia,
does not exhibit pressure reversal. Biophysical Journal 60:1309–1314.
47. Mozhaev, V. V., K. Heremans, J. Frank, P. Masson, and C. Balny, 1996. High pressure effects on protein structure and function.
Proteins 24:81–91.
48. Liu, N.-I., and R. L. Kay, 1977. Redetermination of the pressure dependence of the lipid bilayer phase transition. Biochemistry
16:3484–3486.
49. Winter, R., and C. Jeworrek, 2009. Effect of pressure on membranes. Soft Matter 5:3157–3173.
50. Machta, B. B., S. Papanikolaou, J. P. Sethna, and S. L. Veatch, 2011. Minimal model of plasma membrane heterogeneity requires
coupling cortical actin to criticality. Biophys J 100:1668–77.
51. Zhao, J., J. Wu, and S. L. Veatch, 2013. Adhesion stabilizes robust lipid heterogeneity in supercritical membranes at physiological
temperature. Biophys J 104:825–34.
52. Diaz-Rohrer, B. B., K. R. Levental, K. Simons, and I. Levental, 2014. Membrane raft association is a determinant of plasma membrane
localization. Proceedings of the National Academy of Sciences 111:8500–8505.
53. Lin, Q., and E. London, 2013. Altering Hydrophobic Sequence Lengths Shows That Hydrophobic Mismatch Controls Affinity for
Ordered Lipid Domains (Rafts) in the Multitransmembrane Strand Protein Perfringolysin O. J Biol Chem. 288:1340–1352.
54. Niemela, P. S., S. Ollila, M. T. Hyvonen, M. Karttunen, and I. Vattulainen, 2007. Assessing the Nature of Lipid Raft Membranes.
PLoS Comput Biol 3:1–9.
55. Cicuta, P., S. L. Keller, , and S. L. Veatch, 2007. Diffusion of Liquid Domains in Lipid Bilayer Membranes. The Journal of Physical
Chemistry B 111:3328–3331.
56. Baumgart, T., S. T. Hess, and W. W. Webb, 2003. Imaging coexisting fluid domains in biomembrane models coupling curvature and
line tension. Nature 425:821–824.
57. Franks, N. P., and W. R. Lieb, 1982. Molecular mechanisms of general anaesthesia. Nature 300:487–93.
58. Oleksiuk, O., V. Jakovljevic, N. Vladimirov, R. Carvalho, E. Paster, W. S. Ryu, Y. Meir, N. S. Wingreen, M. Kollmann, and V. Sourjik,
000 00(00) 1–20
20
2011. Thermal Robustness of Signaling in Bacterial Chemotaxis. Cell 145:312–321.
59. Kidd, P. B., M. W. Young, and E. D. Siggia, 2015. Temperature compensation and temperature sensation in the circadian clock.
Proceedings of the National Academy of Sciences 112:E6284–E6292.
60. Daniels, B. C., Y.-J. Chen, J. P. Sethna, R. N. Gutenkunst, and C. R. Myers, 2008. Sloppiness, robustness, and evolvability in systems
biology. Current Opinion in Biotechnology 19:389 – 395.
61. Smith, S. M., R. Renden, and H. von Gersdorff, 2008. Synaptic vesicle endocytosis: fast and slow modes of membrane retrieval.
Trends in Neurosciences 31:559 – 568.
62. Thompson, S. M., L. M. Masukawa, and D. A. Prince, 1985. Temperature dependence of intrinsic membrane properties and synaptic
potentials in hippocampal CA1 neurons in vitro. J. Neurosci. 5:817–24.
63. Tang, L. S., M. L. Goeritz, J. S. Caplan, A. L. Taylor, M. Fisek, and E. Marder, 2010. Precise Temperature Compensation of Phase in
a Rhythmic Motor Pattern. PLoS Biol 8:1–13.
64. Meyer, H., 1901. Zur Theorie der Alkoholnarkose. 3. Mittheilung: Der Einfluss wechselnder Temperatur auf Wirkungsstrke und
Theilungscoefficient der Narcotiea. Arch. exp. Path. Pharmak. (Naunyn-Schmiedebergs) 46:338346.
65. Cherkin, A., and J. F. Catchpool, 1964. Temperature Dependence of Anesthesia in Goldfish. Science 144:1460–1462.
66. Baraska, J., and P. Wlodawer, 1969. Influence of temperature on the composition of fatty acids and on lipogenesis in frog tissues.
Comparative Biochemistry and Physiology 28:553–570.
67. Anderson, T. R., 1970. Temperature adaptation and the phospholipids of membranes in goldfish (carassius auratus). Comparative
Biochemistry and Physiology 33:663–687.
68. Janoff, A. S., M. J. Pringle, and K. W. Miller, 1981. Correlation of General Anesthetic Potency with Solubility in Membranes.
Biochimica Et Biophysica Acta 649:125–128.
69. Bristow, D. R., and I. L. Martin, 1987. Solubilisation of the gamma-aminobutyric acid/benzodiazepine receptor from rat cerebellum:
optimal preservation of the modulatory responses by natural brain lipids. J Neurochem 49:1386–93.
70. Sooksawate, T., and M. Simmonds, 2001. Effects of membrane cholesterol on the sensitivity of the GABAA receptor to GABA in
acutely dissociated rat hippocampal neurones. Neuropharmacology 40:178 – 184.
000 00(00) 1–20
|
1503.06206 | 1 | 1503 | 2015-03-17T14:12:09 | Analysis of the unilateral contact problem for biphasic cartilage layers with an elliptic contact zone and accounting for the tangential displacements | [
"physics.bio-ph",
"physics.med-ph",
"q-bio.TO"
] | A three-dimensional unilateral contact problem for articular cartilage layers attached to subchondral bones shaped as elliptic paraboloids is considered in the framework of the biphasic cartilage model. The main novelty of the study is in accounting not only for the normal (vertical), but also for tangential vertical (horisontal) displacements of the contacting surfaces. Exact general relationships have been established between the contact approach and some integral characteristics of the contact pressure, including the contact force. Asymptotic representations for the contact pressure integral characteristics are obtained in terms of the contact approach and some integral characteristics of the contact zone. The main result is represented by the first-order approximation problem. | physics.bio-ph | physics |
Analysis of the unilateral contact problem for biphasic
cartilage layers with an elliptic contact zone and
accounting for the tangential displacements
A.A. Koroleva∗, S.V. Rogosin†, G.S. Mishuris‡
July 2, 2021
Abstract: A three-dimensional unilateral contact problem for articular cartilage layers at-
tached to subchondral bones shaped as elliptic paraboloids is considered in the framework
of the biphasic cartilage model. The main novelty of the study is in accounting not only
for the normal (vertical), but also for tangential vertical (horisontal) displacements of the
contacting surfaces. Exact general relationships have been established between the contact
approach and some integral characteristics of the contact pressure, including the contact
force. Asymptotic representations for the contact pressure integral characteristics are ob-
tained in terms of the contact approach and some integral characteristics of the contact zone.
The main result is represented by the first-order approximation problem.
1
Introduction
Biomechanical contact problems involving transmission of forces across biological joints are
of considerable practical interest (see, e.g.
[2, 3, 11, 13]). Many analytical solutions to
the problem of contact interaction of articular cartilage surfaces in joints are available. In
particular, Ateshian et al. [8] obtained an asymptotic solution for the axisymmetric contact
problem for two identical biphasic cartilage layers consisting of a solid phase and a fluid
∗Belarusian State University, Belarus.
†Aberystwyth University, UK, and Belarusian State University, Belarus.
‡Aberystwyth University, UK.
The research leading to these results has received funding from the People Programme (Marie Curie
Actions) of the European Union's Seventh Framework Programme FP7/2007-2013/ under REA grant agree-
ment PIRSES-GA-2013-610547 - TAMER and by FP7-PEOPLE-2012-IAPP through the grant PIAP-GA-
2012-284544-PARM2. The authors are thankful to Dr. I. Argatov for important suggestions improving the
proposed model.
1
phase and attached to two rigid impermeable spherical bones of equal radii. Later, Wu
et al.
[14] extended this solution to a more general axisymmetric model by combining the
assumption of the kinetic relationship from classical contact mechanics [12] with the joint
contact model [8] for the contact of two biphasic cartilage layers. An improved solution for
the contact of two biphasic cartilage layers in the axisymmetric setting, which can be used
for dynamic loading, was obtained by Wu et al. [15].
An asymptotic modeling approach to study the contact problem for biphasic cartilage
layers has been performed by Argatov and Mishuris in a series of articles (see [4, 5, 7]). In
particular, it was shown [4] that accounting for the tangential displacements is important in
the case of diseased cartilage where the measurement of indentation depth may differ even
as much as 10% in comparison with the healthy case. In [5], the unilateral contact problem
for articular cartilages bonded to subchondral bones with a contact zone in the shape of
an arbitrary ellipse has been considered, and a closed form analytic solution was found.
Exploiting this exact result, Argatov and Mishuris [7] have performed perturbation analysis
of the contact problem with approximate geometry of the contact surfaces. Other analytic
solutions for the contact problem were found using the viscoelastic cartilage model for elliptic
contact zone in [6]. A new methodology for modeling articular tibio-femoral contact based
on the developed asymptotic model of frictionless elliptical contact interaction between thin
biphasic cartilage layers was presented in [2]. The mathematical model of articular contact
was extended to the case of contact between arbitrary viscoelastic incompressible coating
layers.
In this study we extend results obtained in papers [4] and [5] by considering the influence
of the tangential displacements on the contact problem for cartilage layers with the contact
zone of elliptic shape based on the biphasic material model. Note that the perturbation
method proposed in [7] could be one of the options for the analysis, however, the procedure
is too complex to perform even a few asymptotic steps. Here, employing some technique and
ideas from [4] and [5], we propose another way to construct the asymptotics which utilises the
assumption that the shape of the contact zone is an ellipse at the initial stage of deformation
and can be regarded as a small perturbation of the ellipse at any other stage of deformation.
The paper is organized as follows. The unilateral contact problem formulation and its
linearization are presented in Section 2, where a special case of the contact configuration
with one cartilage layer being plane and rigid is also considered in detail.
In Section 3,
we derive exact general relationships between the contact approach and some integral char-
acteristics of the contact pressure, including the contact force. In Section 3.3, we obtain
asymptotic representations for the contact pressure integral characteristics in terms of the
contact approach and some integral characteristics of the contact zone. The zero-order and
first-order asymptotic approximations for the solution to the contact problem are obtained in
Sections 4.1 and 4.2, respectively. Namely, the first-order approximation problem constitutes
the main result of the present study.
2
2 Formulation of the contact problem
We consider a frictionless contact between two thin linear biphasic cartilage layers firmly
attached to rigid bones shaped like elliptic paraboloids.
In the Cartesian co-ordinates
(x1, x2, z) = (x, z) the equations for the two cartilage surfaces can be written in the form
z = (−1)nΦ(n)(x), n = 1, 2, where
Φ(n)(x) =
x2
1
2R(n)
1
+
x2
2
2R(n)
2
(2.1)
with R(n)
1 , R(n)
2 being the curvature radii of the n-th bone surface at its apex.
In the undeformed state, the cartilage-bone systems occupy convex domains z ≤ −Φ(1)(x)
and z ≥ Φ(2)(x), respectively. They are in the initial contact with the plane z = 0 at the
origin of the co-ordinate system.
We denote by w1(x, t), w2(x, t) the local vertical displacements of the corresponding
cartilage surfaces. Let also u1(x, t), u2(x, t) be the local horizontal (tangential) displacements
of the corresponding surface of the cartilages. Finally, we denote by P (x, t) the contact
pressure density. In this notation the equations for the cartilage surfaces can be written in
the following form:
z = δ1(t) − Φ(1) (x + u1(x, t)) + w1(x, t),
z = −δ2(t) + Φ(2) (x + u2(x, t)) − w2(x, t).
(2.2)
Here, δ1, δ2 are some (positive) vertical displacements of the rigid bones. Note also that
the vertical displacements w1, w2 are positive, while the tangential displacements u1, u2 are
directed outside of the contact zone. Denoting by δ∗(t) = δ1(t) + δ2(t) the contact approach
of the bones, we get from (2.2) the following inequality:
δ∗(t) + w1(x, t) + w2(x, t) ≤ Φ(1) (x + u1(x, t)) + Φ(2) (x + u2(x, t)) .
(2.3)
It was shown in [8] (see also [4]) that vertical and the tangential displacements of each
bone can be represented in the form
wn(x, t) =
∆P (x, t) +
3
Hn
tZ0
hnǫ2
n
3µs,n
un(x, t) = −
hnǫn
2µs,n∇P (x, t), n = 1, 2.
∆P (x, τ )dτ
,
n = 1, 2,
(2.4)
(2.5)
Here ǫn = hn/a0 are dimensionless small parameters, h1, h2 mean the thicknesses of the
cartilage layers, and a0 denotes a characteristic measure of the contact zone (see the detailed
description of the role of this parameter in [4]), Hn = (λs,n + 2µs,n)/µs,n are material pa-
rameters of cartilages, where λs,n and µs,n represent the first Lame coefficient and the shear
3
modulus of the solid phase of the n-th cartilage tissue. Note that u1 and u2 in (2.5) do not
necessarily coincide, they depend on both spatial variables x1, x2, and on the time variable
t.
Following [8], we introduce new spatial variables and time variable via formulas
x′
j =
xj
a0
,
j = 1, 2,
t′ =
χt
µ0
,
where
χ =
3µs,1k1
h2
1
+
3µs,2k2
h2
2
, µ0 =
µs,1
λs,1 + 2µs,1
+
µs,2
λs,2 + 2µs,2
,
a0 is a characteristic measure of the contact zone, and k1, k2 are the cartilage's permeabilities.
In these variables we have the following relations on the contact area ω(t) encircled by the
curve Γ(t) = ∂ω(t):
∆P (x′, t′) + χ
,
(2.6)
h3
2
3µs,2(cid:19)
t′Z0
∆P (x′, τ ′)dτ ′
w1(x′, t′) + w2(x′, t′) =(cid:18) h3
1
3µs,1
+
Φ(n)(x′ + un(x′, t′)) ≃ Φ(n)(x′) −
h2
na0
2µs,n∇Φ(n)(x′) · ∇P (x′, t′), n = 1, 2.
(2.7)
Further the equality in (2.3), i.e.,
δ∗(t′) + w1(x′, t′) + w2(x′, t′) = Φ(1) (x + u1(x, t)) + Φ(2) (x + u2(x, t)) ,
(2.8)
determines the contact area ω(t).
Now we substitute (2.6), (2.7) into (2.8) and obtain the governing equation relating the
contact pressure with the vertical approach of the bones δ∗(t) in the following form (from
now on we keep the names of new unknown functions, e.g. Φ(x) := Φ(x′a0) etc.):
∆P (x, t) + χ
tZ0
∆P (x, τ )dτ = m(cid:16)Φ(x) − δ∗(t) − ∇eΦ(x) · ∇P (x, t)(cid:17) .
Here we have introduced the notation
m =(cid:18) h3
1
3µs,1
+
h3
2
3µs,2(cid:19)−1
,
Φ(x′) = Φ(1)(x′) + Φ(2)(x′).
Thus, it follows from (2.1) and (2.11) that the functions Φ and eΦ are given by
Φ(x) = Φ(x1, x2) = Ax2
1 + Bx2
2
4
(2.9)
(2.10)
(2.11)
(2.12)
with
and
A =
1
2R(1)
1
+
1
2R(2)
1
, B =
1
2R(1)
2
+
1
2R(2)
2
eΦ(x) =eΦ(x1, x2) = eAx2
1 + eBx2
2.
(2.13)
(2.17)
(2.18)
(2.19)
than unit.
Note that the coefficients in eA and eB are positive dimensionless numbers, which are less
Without loss of generality, one can assume that A > B. Then, Eq.
(2.9) can be
rewritten in an equivalent form, using all dimensionless parameters:
∆Pε(x, t) + χ
tZ0
∆Pε(x, τ )dτ = µ(cid:0)Ψ1(x) − δε(t) − ε∇Ψ2(x) · ∇Pε(x, t)(cid:1),
where the following notation has been introduced:
Ψj(x) = x2
1 + e2
j x2
2,
j = 1, 2,
δε(t) =
µ = Am,
It is important to note that
e1 =pB/A,
e2 =qeB/eA,
χ = O(1), µε ≪ χ.
1
A
δ∗(t),
ε = eA
A
.
(2.14)
(2.15)
(2.16)
Discussion of the characteristic values of the introduced parameters is presented in Section
5 (see also [8, 4]).
Since the solution of (2.14) depends on the parameter ε, it is customer to denote an
unknown contact pressure by P = Pε in what follows. Note that the problem for ε = 0
coincides with that considered in [5], where an exact solution to this problem was found.
Equation (2.14) is the equation for determination of the contact pressure Pε(x, t) ≥ 0,
x ∈ ωε(t). In particular, in the case when the contact domain is represented by an ellipse
ωε(t) =(cid:26)x ∈ R2 :
x2
1
b2(t, ε)
+
β2(t, ε)x2
2
b2(t, ε) ≤ 1(cid:27) .
We supply Eq.
(2.14) with the following boundary conditions:
Pε(x, t) = 0, x ∈ Γ(t),
∂Pε
∂n
(x, t) = 0, x ∈ Γ(t).
Note that in the axisymmetric case formula (2.14) coinsides with formula [4, (8)].
5
The equilibrium equation
ZZωε(t)
Pε(x, t)dx = F (t)
(2.20)
connects the external load F (t), unknown contact pressure Pε(x, t), and unknown contact
domain ωε(t).
2.1 Special case of the contact configuration
In order to check the content of formula (2.9) we consider here a special case, namely, we
suppose that the lower part cartilage layer is plane and rigid (the same assumption was
employed in [14]), it means that µs,2 = ∞ and R(1)
2 = ∞, i.e.,
1 = R(1)
Φ(1) ≡ 0, Φ ≡ Φ(2).
In this case we have got the following equation for determination of the contact domain ω(t)
in the form similar to (2.9):
∆P (x, t) + χ
Here we will have
tZ0
∆P (x, τ )dτ = m(cid:16)Φ(x) − δ∗(t) − ∇eΦ(x) · ∇P (x, t)(cid:17) .
m =
3µs,2
h3
2
, χ =
3µs,2k2
h2
2
.
At the same time, small changes have to be made in the right-hand side of Eq.
follows:
Φ(x) =
x2
1
2R(2)
1
h2
2a0x2
1
2µs,2R(2)
1
+
+
,
x2
2
2R(2)
2
h2
2a0x2
1
2µs,2R(2)
1
.
Φ(x) =
Thus Eq.
(2.21) can be rewritten as
∆P (x, t) +
3µs,2k2
h2
2
tZ0
∆P (x, τ )dτ =
2 x2
1
2R(2)
1
3µs,2
h3
+
x2
2
2R(2)
2
− δ∗(t)!
3a0
h2 " x1
R(2)
1
−
∂x1P (x, t) +
x2
R(2)
2
∂x1P (x, t)# .
(2.21)
(2.22)
(2.21) as
(2.23)
(2.24)
It can be easily checked that in the axisymmetric case Eq. (2.23) reduces to the governing
differential equation obtained in [4].
6
3 General relationships between the solution compo-
nents
3.1 Determination of the contact approach
In our model we assume that the external load is non-decreasing. Thus, the contact domain
is monotonically expanded, i.e.
ωε(t1) ⊆ ωε(t2),
∀t1 ≤ t2.
(3.1)
It is convenient to suppose also that the contact pressure is defined on the whole plane. For
this we simply extend the density Pε(x, t) by assuming that
Integrating (2.14) over contact domain ω(t), we get
Pε(x, t) = 0,
∀x 6∈ ωε(t).
ZZω(t)
∆Pε(x, t)dx + χZZω(t)
tZ0
∆Pε(x, τ )dτ dx =
= µZZω(t)
(Ψ1(x) − δε(t)) dx − εµZZω(t)
∇Ψ2(x) · ∇Pε(x, t)dx.
(3.2)
(3.3)
For simplicity of notation, we omit here (and everywhere in the next two sections) the
subindex ε in ωε.
From the monotonicity of the contact domain (3.1) and assumption (3.2), it follows that
the second integral on the left-hand side can be written in the form
ZZω(t)
tZ0
∆Pε(x, τ )dτ dx =
tZ0 ZZω(t)
∆Pε(x, τ )dxdτ.
(3.4)
Using the second Green's formula
ZZω(t)
(u(x)∆v(x) − v(x)∆u(x)) dx = ZΓ(t) (cid:18)u(x)
∂v
∂n
(x) − v(x)
∂u
∂n
(x)(cid:19) ds
(3.5)
with u ≡ 1 and v = Pε(x, t) we get the following relation in view of the boundary condition
(2.19):
ZZω(t)
∆Pε(x, τ )dx = ZΓ(t)
∂Pε
∂n
(x, s)ds = 0, ∀τ ≤ t.
(3.6)
Therefore, the both integrals on the left-hand side of (3.3) vanish.
7
Further, we use the first Green's formula
ZZω(t)
(ϕ∆ψ + ∇ϕ · ∇ψ) dx = ZΓ(t)
ϕ
∂ψ
∂n
ds
(3.7)
with ψ(x) = Ψ2(x) and ϕ(x) = Pε(x, t). In this case the integral on the right-hand side
vanishes in view of (2.18), and we obtain the relation
ZZω(t)
∇Ψ2(x) · ∇Pε(x, t)dx = −ZZω(t)
Pε(x, t)∆Ψ2(x)dx = −2(1 + e2
2)F (t),
(3.8)
where we used the equilibrium equation (2.20) and the identity
∆Ψ2(x) = 2(1 + e2
2)
(3.9)
with e2 being defined in (2.15).
In what follows, it is convenient to have the following notation for the integrals of the
product of k-th power of the function Ψ1 and l-th power of the function Ψ2:
Ak,l(ω) =ZZω
Ψk
1(x)Ψl
2(x)dx > 0, k, l = 0, 1, 2, . . .
(3.10)
In particular, A0,0(ω) is the area of the contact domain. It is to remember that the constants
Ak,l(ω) depend finally on t, but we omitted this fact in the notation in order to avoid
cumbersome expressions. Computations of Ak,l(ω) for the elliptic domain (2.17) we included
into Appendix (see Section 6.1).
Taking into account Eqs. (3.6) and (3.8), we get
δε(t) =
A1,0(ωε(t))
A0,0(ωε(t))
+
2(1 + e2
2)ε
A0,0(ωε(t))
F (t).
(3.11)
This formula allows us to compute the contact approach δε(t) as a function of the total
external force F (t) and the main axes of the ellipse describing the shape of the contact zone,
which in fact depends on time too.
3.2 Some integral identity for the contact pressure
In order to write out a more informative equation for the contact load, we use the following
trick. We multiply both sides of (2.14) by the function v(x) = Ψ2(x) and integrate the
8
obtained equation over the contact domain ω(t)
ZZω(t)
Ψ2(x)∆Pε(x, t)dx + χZZω(t)
= µZZω(t)
− µεZZω(t)
Ψ2(x)∆Pε(x, τ )dτ dx =
tZ0
Ψ2(x)Ψ1(x)dx − µδε(t)ZZω(t)
Ψ2(x)dx
Ψ2(x)∇Ψ2(x) · ∇Pε(x, t)dx.
(3.12)
Let us calculate the integrals in this relation by using Green's formulas. For the first integral
on the left-hand side we use formula (3.5) with u = Ψ2, v = Pε and the boundary conditions
(2.18), (2.19). Hence, we obtain
ZZω(t)
Ψ2(x)∆Pε(x, t)dx =ZZω(t)
∆Ψ2(x)Pε(x, t)dx.
Now taking into account (3.9), we get
ZZω(t)
Ψ2(x)∆Pε(x, t)dx = 2(1 + e2
2)F (t).
(3.13)
For the second integral on the left-hand side, we apply the same approach, but interchange
first the integrals over ωε(t) and over τ ∈ (0, t) exploiting the load monotonicity. Therefore,
we arrive at the equation
ZZω(t)
tZ0
Ψ2(x)∆Pε(x, τ )dτ dx =
tZ0 ZZω(t)
Ψ2(x)∆Pε(x, τ )dτ dx = 2(1 + e2
2)
tZ0
F (τ )dτ.
(3.14)
For the first and second integrals on the right-hand side, we simply use the notation (3.10),
which gives
ZZω(t)
Ψ1(x)Ψ2(x)dx = A1,1(b; β),
ZZω(t)
Ψ2(x)dx = A0,1(b; β).
(3.15)
Finally, for the third integral on the right-hand side, we make use of the following simple
formula which follows immediately from the definition of Ψ2:
Ψ2∇Ψ2 =
1
2∇Ψ2
2.
9
Then we can apply Green's formula (3.7) and the boundary conditions (2.18), (2.19) to find
ZZω(t)
Ψ2(x)∇Ψ2(x) · ∇Pε(x, t)dx = −
1
2ZZω(t)
∆Ψ2
2(x)Pε(x, t)dx.
By applying the second Green's formula (3.5) with u = Pε, v = Ψ2
conditions (2.18), (2.19), we represent this integral in the form
2, and the boundary
ZZω(t)
Ψ2(x)∇Ψ2(x) · ∇Pε(x, t)dx = −
1
2ZZω(t)
Ψ2
2(x)∆Pε(x, t)dx.
(3.16)
This integral still contains the unknown density of contact pressure Pε(x, t). Let us define
M(j)Pε(t) ≡ZZω(t)
Ψj
2(x)∆Pε(x, t)dx.
(3.17)
Now we rewrite the relation (3.12) by using the results for all integrals (3.13) -- (3.16) in
the following form:
2(1 + e2
2)KF (t) = µA1,1(ωε(t)) − µδε(t)A0,1(ωε(t)) +
µε
2 M(2)Pε(t).
(3.18)
Here, we have introduced the Volterra operator K as follows:
KF (t) = F (t) + χ
F (τ )dτ.
(3.19)
tZ0
Note that the integral in the right-hand side of the equation (3.18) allows to continue
the same procedure to deliver an asymptotic estimate for this equation.
We continue to proceed with Eq.
(3.18) on the next steps.
3.3 Asymptotic estimates of the integral characteristics M(j)Pε(t)
Now we proceed to calculate the last integral in (3.18). For this we multiply the governing
integral equation (2.14) by Ψj
2(x) (j ≥ 2) and integrate over contact domain ω(t):
ZZω(t)
Ψj
2(x)∆Pε(x, t)dx + χZZω(t)
= µZZω(t)
− µεZZω(t)
Ψj
10
Ψj
2(x)∆Pε(x, τ )dτ dx =
tZ0
2(x)Ψ1(x)dx − µδε(t)ZZω(t)
Ψj
2(x)dx
Ψj
2(x)∇Ψ2(x) · ∇Pε(x, t)dx.
(3.20)
By using the same argument as on the previous step, we get
KM(j)Pε(t) = µA1,j − µδε(t)A0,j(a; β) − µεZZω(t)
Ψj
2(x)∇Ψ2(x) · ∇Pε(x, t)dx.
(3.21)
For the last integral we use the relations
Ψj
2(x)∇Ψ2(x) =
1
j + 1∇Ψj+1
2
(x)
and
ZZω(t)
∇Ψj+1
2
(x) · ∇Pε(x, t)dx = −ZZω(t)
∆Ψj+1
2
(x)Pε(x, t)dx.
Therefore, the integral
M(j)Pε(t) = µK−1(cid:26)A1,j(ωε(t)) − δε(t)A0,j(ωε(t)) +
ε
j + 1KM(j+1)Pε(t)(cid:27)
(3.22)
has been obtained as a solution of the integral equation (3.21). Here the inverse operator
K−1 is defined by the formula
K−1Y (t) = Y (t) − χ
tZ0
Y (τ )e−χ(t−τ )dτ.
(3.23)
Performing the same computation, we obtain the following representation for the integral
in the right-hand side of (3.18):
M(2)Pε(t) =
2εj−1
(j + 1)!
NXj=1
µjK−j {A1,j+1(ωε(t)) − δε(t)A0,j+1(ωε(t))}
+
2εN
(N + 2)!
Substituting this representation into Eq.
µNK−NM(N +2)Pε(t).
(3.18), we finally get
2(1 + e2
2)KF (t) =
NXj=0
+
εj
(j + 1)!
µj+1K−j {A1,j+1(ωε(t)) − δε(t)A0,j+1(ωε(t))}
εN +1
(N + 2)!
µN +1K−NM(N +2)Pε(t),
or equivalently
2(1 + e2
2)KN +1F (t) =
εj
(j + 1)!
NXj=0
µj+1KN −j {A1,j+1(ωε(t)) − δε(t)A0,j+1(ωε(t))}
11
(3.24)
(3.25)
+
εN +1
(N + 2)!
µN +1M(N +2)Pε(t).
(3.26)
The latter relation allows us to determine the problem parameters asymptotically with any
prescribed accuracy.
Note that apart from the fact that the shapes of the contacting bones are elliptical
paraboloids, no additional assumptions on the shape of the contact zone have been made.
On the other hand, no proof was offered to show that the contact zone is approximately
represented by an ellipse. This will be done later.
Remark 1. For every t for which the contact pressure Pε(t) is bounded and the contact
region ω(t) belongs to a bounded domain, the remainder εN+1
(N +2)! µN +1M(N +2)Pε(t) in formula
(3.26) tends to zero as N → ∞. Thus, the series corresponding to the sum on the right
hand-side of (3.26) is converging.
4 Asymptotic solution to the contact problem
4.1 Zero-order approximation
First, we get solution of the problem for ε = 0. In this case Eq.
(2.14) has the form
∆P (0)(x, t) + χ
tZ0
∆P (0)(x, τ )dτ = µ(cid:0)Ψ1(x) − δ(0)(t)(cid:1) ,
(4.1)
where Ψ1(x) is defined in (2.15). Since we know from [5] that the contact zone is an ellipse
at this stage of approximation we will have
δε = δ(0)(t) = δε(b0(t); β0(t)) =
A1,0(ω0(t))
A0,0(ω0(t))
.
(4.2)
Using formula (4.2) and calculations presented in Section 6.1 (see formula (6.6)), one
can find that
and therefore
A0,0(ω0(t)) =
πb2
0
β0
, A1,0(ω0(t)) =
πb4
0
4β3
0(cid:0)β2
0 + e2
1(cid:1) ,
δ(0)(t) =
b2
0 (β2
0 + e2
1)
4β2
0
.
Note that formulas (4.3) and (4.4) contain two known constants e1 and e2 defined in (2.15)
and two still unknown functions b0(t) and β0(t), which are the main semi-axis and the
eccentricity of the ellipse
ω0(t) =(cid:26)x ∈ R2 :
β2
0(t)x2
2
b2
0(t) ≤ 1(cid:27) .
+
x2
1
b2
0(t)
12
(4.3)
(4.4)
(4.5)
πb6
0
24β5
0(cid:8)3β4
0(cid:8)3β4
πb6
0
48β5
2(cid:9) ,
2(cid:9) .
1e2
2(1 + e2
2)KF (t) = µ
0 − (e2
1 + e2
2)β2
0 + 3e2
To find the functions b0(t) and β0(t) together with the pressure distribution over the
contact zone, P (0)(x, t), we follow [5] and introduce a new unknown function
p(0)(x, t) = P (0)(x, t) + χ
tZ0
P (0)(x, τ )dτ = KP (0)(x, t).
(4.10)
In the case of monotone external load, this function should satisfy the Poisson equation
(following from (2.9))
∆p(0)(x, t) = µ(cid:0)Ψ1(x) − δ(0)(t)(cid:1) , x ∈ ω0(t),
with the boundary conditions (2.18), (2.19).
It is customary to rewrite this relation in the form
G0(x, t) = 0,
where
(4.7)
(4.8)
(4.9)
(4.11)
(4.12)
(4.13)
(4.14)
The leading terms in (3.26) imply (for N = 0) the following equation:
2(1 + e2
2)KF (t) = µA1,1(ω0(t)) − µδ(0)(t)A0,1(ω0(t)).
(4.6)
Here, K is the Volterra integral operator defined in (3.19).
Analogously, using some results from Section 6.1 (see, in particular, formula (6.6)), we
obtain
and
and thus
A0,1(ω0(t)) =
0 + e2
πb4
0
4β3
0 (cid:0)β2
2(cid:1)
A1,1(ω0(t)) =
0 + (e2
1 + e2
2)β2
0 + 3e2
1e2
G0(x, t) = G0(b0, β0, δ0)
≡ ∆p(0)(x, t) − µ(cid:0)Ψ1(x) − δ(0)(t)(cid:1) , x ∈ ω0(t).
Bearing in mind that the function Ψ1(x) is a quadratic polynomial (compare with (2.15)),
it is natural to look for the solution of such problem in the form of a polynomial in x1, x2 of
the fourth degree, that is
p(0)(b0, β0, η0, x, t) = η0(t)(cid:18)1 −
x2
1
0 −
b2
β2
0x2
2
b2
0 (cid:19) Q0(x1, x2).
(4.15)
13
Note that the term in the brackets vanishes on the boundary ω0, and thus the condition
(2.18) is satisfied automatically.
In Section 6.2, it has been shown that Q0 is a polynomial of the second order having the
form
so that
x2
1
0 −
b2
Q0(x1, x2) =(cid:18)1 −
p(0)(x1, x2; t) = η0(t)(cid:18)1 −
β2
0x2
2
b2
0 (cid:19) ,
0 (cid:19)2
x2
1
0 −
b2
β2
0x2
2
b2
(4.16)
(4.17)
.
Taken into account this representation we arrive at the following relations (see Section 6.3):
η0(t) =
µδ(0)(t)
4(1 + β2
0)
b2
0,
η0(t) =
η0(t) =
µb4
0
2(6 + 2β2
0)
µb4
0 + 6β4
0)
0e2
1
2(2β2
=
=
µb4
0
,
4(3 + β2
0)
0e2
1
µb4
0(1 + 3β2
0)
4β2
(4.18)
(4.19)
(4.20)
.
This system allows us to determine the unknown functions b0(t) and β0(t).
Indeed,
eliminating η0 from the last two equations, we get a bi-quadratic equation defining the value
of the parameter β0, i.e.,
By definition, β0 is a positive parameter, thus the unique positive solution of (4.21) has the
form
3β4
0 + (1 − e2
1)β2
0 − 3e2
1 = 0.
(4.21)
β0 =s (e2
1 − 1) +pe4
6
1 + 34e2
1 + 1
.
(4.22)
Note that at the zero-approximation the parameter β0 does not depend on time. The other
parameter, η0(t), can be computed directly from (4.19) or (4.20), if one knows the remaining
constant b0(t). Moreover, taking into account (4.18) and (4.4), one can use an equivalent
formula
η0(t) =
µb4
16β2
0 + e2
0(β2
1)
0(1 + β2
0)
.
(4.23)
In the same way, one can offer, in addition to (4.4), two equivalent representations for
the indentation parameter
δ(0)(t) =
1 + β2
0
3 + β2
0
b2
0(t) =
0)e2
(1 + β2
1
0(1 + 3β2
β2
0)
b2
0(t).
(4.24)
Finally, the major semi-axis b0 of the ellipse ω0 is determined as follows:
b0(t) =F (t) + χ
tZ0
F (τ )dτ(cid:18)
µπ(3β4
96β5
0 − β2
0(1 + e2
2)
2) + 3e2
1 + e2
0(e2
1e2
14
1/6
2)(cid:19)
.
(4.25)
Note that the parameters b0, η0 as well as the indentation, δ0, depend on time t in contrast
to the ellipse eccentricity β0.
Now, it remains only to find the pressure over the contact area. Using (4.10) and (4.17),
we get
P (0)(b0, β0, η0, x1, x2, t) = K−1(cid:0)η0(t)Q0(x1, x2)2(cid:1) .
0(t) − β2
b2
1
x2
1
0(t) −
b2
If (x1, x2) belongs to the initial contact zone, i.e. 1 − x2
P (0)(x1, x2, t) = η0(t)(cid:18)1 −
η0(τ )(cid:18)1 −
0(t) − β2
If (x1, x2) lies outside of the initial contact zone, i.e. 1 − x2
P (0)(x1, x2, t) = η0(t)(cid:18)1 −
η0(τ )(cid:18)1 −
0(t)(cid:19)2
β2
0 x2
2
b2
0(t)(cid:19)2
x2
1
0(t) −
b2
tZt∗(x1,x2)
0x2
β2
2
b2
tZ0
− χ
−χ
b2
1
x2
1
0(τ ) −
b2
x2
1
0(τ ) −
b2
0 x2
0(t) > 0, then
2
b2
β2
0x2
2
b2
0(τ )(cid:19)2
e−χ(t−τ )dτ.
(4.27)
0x2
0(t) < 0, then
2
b2
(4.26)
0x2
β2
2
b2
0(τ )(cid:19)2
e−χ(t−τ )dτ.
(4.28)
The critical moment of time t∗ is determined by the formula
b2
0(t∗) = x2
1 + β2
0x2
2.
Using (4.25), we get
F (t∗) + χ
t∗Z0
F (τ )dτ =
µπ
96β5
0(cid:18)3β4
0 − β2
0(e2
1 + e2
1 + e2
2
2) + 3e2
1e2
2
(cid:19) (x2
1 + β2
0x2
2)3.
(4.29)
If the load is stepwise, we have F (t) = F0. Hence, we find that
t∗ =
µπ
96β5
0χF0(cid:20)(3β4
0 − β2
0(e2
1 + e2
1 + e2
2
2) + 3e2
1e2
2)
(x2
1 + β2
0x2
2)3(cid:21) −
1
χ
.
(4.30)
Note that in this case
b6
0(t∗) =
96β5
0(1 + e2
0 − β2
0(e2
µπ(3β4
2)(1 + χt∗)
1 + e2
2) + 3e2
1e2
2)
F0.
(4.31)
This finishes the zero iteration step. Note that the results of this Section after changing
the notation coincide with those obtained in [4].
15
4.2 First-order approximation problem
For the next steps we consider an appropriately deformed contact domain ω(1)
ε , defined as a
perturbation of the zero-order one ω0. Namely, we assume that it can be written in the form
ω(1)
ε = ω(1)
ε (t) =n(x1, x2) : Q0(x, t) + εQ1(x, t) ≥ 0o,
where unknown polynomials are taken in the forms
Q0(x, t) = Q0(x, β1, b1),
Q1(x, t) = a40(t)x4
1 + a22(t)x2
1x2
2 + a04(t)x4
2.
(4.32)
(4.33)
(4.34)
Note that for ε = 0 the solution form coincides with (4.5), if one take b1 ≡ b0, β1 ≡ β0.
The idea behind such choice of the asymptotic anzatz is to satisfy the boundary condi-
tions (2.18) and (2.19) automatically. This will be archived by putting
P (1)
ε = K−1(cid:16)η(1)(t)(cid:0)Q0(x1, x2, β1(t), b1(t)) + εQ1(x, t)(cid:1)2(cid:17).
Now, when the boundary conditions are valid, we will satisfy the governing equation
(2.9). Note that
P (1)
ε = P0 + εP1 + O(ε2),
where pj = K(Pj), j = 0, 1, and
p0 = η(1)(t)(cid:18)1 −
p1 = 2η(1)(t)(cid:18)1 −
K(cid:0)∆(P (0) + εP1 + O(ε2))(cid:1) = µ(cid:0)Ψ1 − δ(1)
Substituting this representation into Eq.
,
β2
1(t)x2
2
b2
x2
1
1(t) −
b2
β2
1(t)x2
2
b2
1(t) (cid:19)2
1(t) (cid:19) Q1(x, t).
x2
1
1(t) −
b2
(2.9), we obtain
where the parameter δ(1)
ε
is represented in the same form as P (1)
ε
, i.e.,
ε − ε∇Ψ2 · (∇P (0) + ε∇P (1) + O(ε2))(cid:1) ,
(4.35)
(4.36)
(4.37)
(4.38)
(4.39)
(4.40)
(4.41)
δ(1)
ε = δ0 + εδ1 + O(ε2) = δ(1) + O(ε2).
We can write Eq.
(4.39) with the accuracy to the terms of O(ε2) as follows:
∆p(0) + ε∆p1 = µ(cid:0)Ψ1 − δ(1) − ε∇Ψ2 · ∇P (0)(cid:1) .
An extended variant of this equation can be written by using the definition of all com-
ponents of the equation and by comparing coefficients at different powers of x1, x2, so that
16
4η(1)
b2
1
−
(1 + β2
1) = −µδ(1),
1
4η(1)(cid:20)3 + β2
4η(1)(cid:20) β2
1(1 + 3β2
1)
+ ε(6a40 + a22)(cid:21) = µ(1 − 8εθ2,0),
+ ε(a22 + 6a04)(cid:21) = µ(e2
1 − 8εe2
b4
1
b4
1
2θ2,2),
24η(1)
b2
1
− ε
(a40β2
1 + a22(1 + β2
1) + a04) = 8εµ(1 + e2
2)θ4,2,
4η(1)
b2
1
− ε
4η(1)
− ε
b2
1
(a40(15 + β2
1) + a22) = 8εµθ4,0,
(a04(15β2
1 + 1) + a22β2
1) = 8εµe2
2θ4,4,
(4.42)
(4.43)
(4.44)
(4.45)
(4.46)
(4.47)
(4.48)
where
θ2k,2l(t) = K−1(cid:0)η(1)b−2k
1 β2l
1(cid:1) , k, l = 0, 1, 2.
In the system (4.42) -- (4.47) we have 6 equations and 7 unknowns: η(1)(t), δ(1)
ε
, b1(t), β1(t),
and a40, a22, a04 (coefficients of the polynomial Q1). Therefore, we have to add an extra
equation to the above system, namely
δ(1)(t) =
A1,0(ωε(t))
A0,0(ωε(t))
+
2(1 + e2
2)ε
A0,0(ωε(t))
F1(t),
(4.49)
where F1(t) can be represented in the form
F1(t) =ZZω(1)
ε
P (1)
ε (x, t)dx.
We also make use of Eq.
O(ε2) in the form
(3.26) written for this approximation step with the accuracy of
2(1 + e2
2)K2F (t) =
1Xj=0
εj
(j + 1)!
µj+1K1−j(cid:8)A1,j+1(ωε(t)) − δ(1)(t)A0,j+1(ωε(t))(cid:9) .
(4.50)
Remark 2. Note that putting ε = 0, the system (4.42) -- (4.47), (4.49) transforms to the
previous case evaluated in the previous section.
Remark 3. In the case when ε > 0, the system (4.42) -- (4.47), (4.49) has to be solved
numerically. Note that the parameter ε in the last three equations (4.45) -- (4.47) can be
canceled. We left these multipliers here to explain the limiting case (ε = 0).
17
5 Discussion and conclusion
First of all, observe that at t = 0, the contact problem for biphasic layers reduces to that for
elastic incompressible layers. The contact problem in the latter case were studied in a number
of papers [1, 9, 10, 16], however, without taking into account the tangential displacements.
To solve the resulting problem (4.42) -- (4.47) and (4.49), we suggest the following iterative
algorithm:
• Taking ε = 0, we have computed all values η, b, β, δ = η0, b0, β0, δ0 from the zero-order
approximation.
• Having them we can compute the quantity θ2k,2l(t) from (4.48),
• Then, from the system of three equations (4.45) -- (4.47) we compute the constants
a40, a22, a04 assuming the values of η, b, β as above.
• Finally from the system of four equations (4.42) -- (4.44) and (4.49) considering the right-
hand side known (computed by the values know from the previous computations), we
found new values η, b, β, δ and compare them with the previous computations. If the
required accuracy has achieved we stop the computation, if not we are going to the
second step of this iterative procedure.
We note that formulas (2.4) and (2.5) for the vertical and tangential displacements
contain different powers of parameters ǫ, namely, ǫ2 and ǫ, respectively. Note also that our
analysis (with the values of another parameters taken into account) shows, that the role
of these magnitudes (vertical and tangential displacements) is quite opposite. In the final
equation (see (2.14)) the leading terms, corresponding to the vertical displacement, contain
the zero power of the new small parameter ε, but the leading terms, corresponding to the
tangential displacements, contain the first power of ε.
References
[1] Aleksandrov, V.M. Asymptotic solution of the axisymmetric contact problem for an
elastic layer of incompressible material. J. Appl. Math. Mech. 67, 589 -- 593 (2003)
[2] Argatov, I. Development of an asymptotic modeling methodology for tibio-femoral con-
tact in multibody dynamic simulations of the human knee joint. Multibody Syst. Dyn.
28, 3 -- 20 (2012).
[3] Argatov, I. Contact problem for a thin elastic layer with variable thickness: Application
to sensitivity analysis of articular contact mechanics. Appl. Math. Model. 37, 8383 -- 8393
(2013).
18
[4] Argatov, I., Mishuris, G. Axisymmetric contact problem for a biphasic cartilage layer
with allowance for tangential displacements on the contact surface. Eur. J. Mech.
A/Solids 29, 1051 -- 1064 (2010).
[5] Argatov, I., Mishuris, G. Elliptical contact of thin biphasic cartilage layer: Exact solu-
tion for monotonic loading. J. Biomech. 44, 759 -- 761 (2011).
[6] Argatov, I., Mishuris, G. Frictionless elliptical contact of thin viscoelastic layers bonded
to rigid substrates. Applied Math. Model. 35, 3201 -- 3212 (2011).
[7] Argatov, I., Mishuris, G. Contact problem for thin biphasic cartilage layers: perturba-
tion solution. Quart. J. Mech. Appl. Math. 64, 297 -- 318 (2011).
[8] Ateshian, G.A., Lai, W.M., Zhu, W.B., Mow, V.C. Anasymptotic solution for the con-
tact of two biphasic cartilage layers. J. Biomech. 27, 1347 -- 1360 (1994).
[9] Barber, J.R.: Contact problems for the thin elastic layer. Int. J. Mech. Sci. 32, 129 -- 132
(1990).
[10] Chadwick, R.S. Axisymmetric indentation of a thin incompressible elastic layer. SIAM
J. Appl. Math. 62, 1520 -- 1530 (2002).
[11] Hunziker, E. B. Articular cartilage repair: basic science and clinical progress. A review
of the current status and prospects. Osteoarthritis and Cartilage. 10, 432 -- 463 (2001).
[12] Johnson, K. L. Contact Mechanics. Cambridge, Cambridge University Press (1985).
[13] Owen, J. R., Wayne, J. S. Contact models of repaired articular surfaces:
influence of
loading conditions and the superficial tangential zone. Biomech Model Mechanobiol. 10,
461 -- 471 (2011).
[14] Wu, J. Z., Herzog, W., and Ronsky, J. Modeling axi-symmetrical joint contact with
biphasic cartilage layers -- An asymptotic solution. J. Biomech. 29, 1263 -- 1281 (1996).
[15] Wu, J. Z., Herzog, W., and Epstein, M. An improved solution for the contact of two
biphasic cartilage layers. J. Biomech. 30, 371 -- 375 (1997).
[16] Yang, F. Indentation of an incompressible elastic film. Mech. Mater. 30, 275 -- 286 (1998).
19
6 Appendix
6.1 Calculation of the constants Akl
Here, we compute the values of the constants
Ak,l(b; β) =ZZω(t)
Ψk
1(x)Ψl
2(x)dx > 0, k, l = 0, 1, 2, . . .
First of all, we note that an unknown contact domain ω(t) is of the same type as sections of
the initial gap elliptical paraboloid, i.e., it is an ellipse coaxial to the ellipse
ω(t) = ωε(t) =(cid:26)x ∈ R2 :
x2
1
b2(t; ε)
+
x2
2β2(t; ε)
b2(t; ε) ≤ 1(cid:27) .
In order to avoid long formulas, we use the short notation for ω(t), writing all parameters
without variables they depend on, i.e.,
ω(t) =(cid:26)x ∈ R2 :
x2
1
b2 +
2β2
x2
b2 ≤ 1(cid:27) .
Performing the standard change of variables
x1 = br cos θ,
x2 =
b
β
sin θ,
we represent the integral for Ak,l(t) in the form
Ak,l(b; β) =
1Z0
2πZ0 (cid:18)b2r2 cos2 θ +
b2e2
2
β2 r2 sin2 θ(cid:19)l b2
β
b2e2
1
β2 r2 sin2 θ(cid:19)k(cid:18)b2r2 cos2 θ +
2πZ0
kXi=0
i!(k − i)!
k!
e2i
1
β2i sin2i θ cos2k−2i θ
r2k+2l+1dr
=
b2k+2l+2
β
1Z0
=
×
lXj=0
l!
j!(l − j)!
b2k+2l+2
(2k + 2l + 2)β
×
lXj=0
l!
j!(l − j)!
e2j
2
β2j sin2j θ cos2l−2j θdθ
2πZ0
kXi=0
k!
i!(k − i)!
e2i
1
β2i sin2i θ cos2k−2i θ
e2j
2
β2j sin2j θ cos2l−2j θdθ.
20
rdrdθ
(6.1)
(6.2)
Since the trigonometric functions are presented here only in even powers, then the last
integration can be performed over the interval [0, π/2] as follows:
Ak,l(b; β) =
4b2k+2l+2
(2k + 2l + 2)β
π/2Z0
kXi=0
k!
i!(k − i)!
e2i
1
β2i sin2i θ cos2k−2i θ
×
lXj=0
l!
j!(l − j)!
e2j
2
β2j sin2j θ cos2l−2j θdθ
=
4b2k+2l+2
(2k + 2l + 2)β
×
lXj=0
l!
j!(l − j)!
kXi=0
e2j
2
β2j
k!
i!(k − i)!
e2i
1
β2i
π/2Z0
sin2i+2j θ cos2k−2i+2l−2j θdθ.
(6.3)
The integrals in (6.3) are calculated by using formulas
π/2Z0
sin2p θ cos2q θdθ =
1
2
Γ(p + 1/2)Γ(q + 1/2)
Γ(p + q + 1)
,
p, q > 0,
(6.4)
and Legendre's duplication formula for the Gamma-function
Γ(n + 1/2) =
√2πΓ(2n)
22n−1/2Γ(n)
, n ∈ N,
(6.5)
as well as the relation Γ(n + 1) = n!. Finally, we arrive at the following representation of
Ak,l = Ak,l(b; β) valid for all k, l ∈ N0 = N ∪ {0}:
Ak,l =
2πb2 (b/2)2k+2l
β(2k + 2l + 2)(k + l)!
kXi=0
k!
i!(k − i)!
e2i
1
β2i
×
lXj=0
l!
j!(l − j)!
e2j
2
β2j
(2i + 2j)!(2k − 2i + 2l − 2j)!
(i + j)!(k − i + l − j)!
(6.6)
.
6.2 Computation of the polynomial Q0
In order to determine the coefficients of the polynomial
Q0(x1, x2) = 1 + q1,0x1 + q0,1x2 + q2,0x2
1 + q1,1x1x2 + q0,2x2
2,
we need to compute the normal derivative of the unknown functions p(0) (4.15) along the
elliptic boundary Γ:
∂p(0)
∂n Γ = ∇p(0) · −→n Γ = η0(t)(cid:18)−
2x2
1
0 −
b2
2β4
0x2
2
b2
0 (cid:19) Q0Γ = 0.
21
(6.7)
Here we take into account the fact that, since the contact domain is an ellipse (4.5), the
tangential and normal vectors to the boundary Γ = ∂Ω are given by
Then, to satisfy the boundary condition (2.19) the following equation should be valid:
−→r =(cid:0)−β2
0x2, x1(cid:1) , −→n =(cid:0)x1, β2
0x2(cid:1) .
(6.8)
(6.9)
(6.10)
,
(6.11)
(6.12)
(6.13)
(6.14)
(6.15)
(6.16)
This, in turn, is equivalent to the representation
Q0Γ = 0.
Q0(x1, x2) =(cid:18)1 −
x2
1
0 −
b2
β2
0x2
2
b2
0 (cid:19) .
6.3 Evaluation of the ellipse parameters
Since
we have
Therefore, by straightforward computations, we find that
β2
0x2
2
b2
0 (cid:19)2
0 (cid:19) ,
0 (cid:21) .
2x1
b2
p(0)(x, t) = p(0)(x1, x2, t) = η0(t)(cid:18)1 −
x2
1
0 −
b2
∂p(0)
∂x1
∂2p(0)
∂x2
1
x2
1
0 −
b2
x2
1
0 −
b2
2
b2
2
b2
0
∂2p(0)
∂x2
1
= 2η0(t)(cid:18)1 −
0(cid:18)1 −
= 2η0(cid:20)−
= 2η0(cid:20)−
= 2η0(cid:18)1 −
x2
1
0 −
b2
0 (cid:18)1 −
x2
1
0 −
b2
= 2η0(t)(cid:20)−
2β2
0
b2
0
2β2
0
b2
+
∂p(0)
∂x2
= 2η0(cid:20)−
β2
0x2
2
b2
β2
0x2
2
b2
2x1
b2
2x1
b2
0
0 (cid:19) ·(cid:18)−
0 (cid:19) +
0 (cid:21) .
0 (cid:19) ,
2β2
0x2
2
b4
+
0x2
β2
2
b2
β2
0x2
2
b2
0 (cid:19) ·(cid:18)−
0 (cid:19) +
2β2
0x2
b2
2β2
0x2
b2
0
6x2
1
b4
0
+
0x2
6β4
2
b4
0 (cid:21) .
2β2
0x2
b2
0 (cid:21) .
∂2p(0)
∂x2
2
Thus, we obtain
0x2
2β2
1
b4
0
Substituting (6.13) and (6.15) into the main equation
∂2p(0)
∂x2
2
+
where
G0 = 2η0(t)(cid:20)(−2)
G0(b0, β0, δ0) ≡ ∆p(0)(x, t) − µ(cid:0)Ψ1(x) − δ(0)(t)(cid:1) = 0,
1 + β2
0
b2
0
+(cid:18) 6 + 2β2
0 (cid:19) x2
1 +(cid:18) 6β4
b4
0
b4
0
0 + 2β2
0
22
2(cid:21) − µ(cid:0)Ψ1(x1, x2) − δ(0)(t)(cid:1) ,
(cid:19) x2
and taking into account that
Ψ1(x) = Ψ1(x1, x2) = x2
1 + e2
1x2
2,
one concludes that the expression for G0 is represented by a second order polynomial with
respect to the independent variables x1 and x2 in the following form:
G0(b0, β0, η0, δ(0)) = q0(b0, β0, η0, δ(0)) + q1(b0, β0, η0)x2
1 + q2(b0, β0, η0)x2
2.
(6.17)
Here the coefficients are defined as follows:
q0(b0, β0, η0, δ(0)) =
4η0
µb2
0
(1 + β2
0) − δ(0),
q1(b0, β0, η0) =
(3 + β2
0) − µ,
q2(b0, β0, η0) =
(1 + 3β2
0) − µe2
1.
4η0
b4
0
4η0β2
0
b4
0
6.4 Auxiliary computation
Taking into account (4.37), we can represent p0(x, t) in the form
p0(x, t) = η(1)(t)(cid:18)1 −
2x2
1
1 −
b2
1x2
2β2
2
b2
1
+
Hence, applying the Laplace equation, we get
2β2
1x2
2
1x2
b4
1
+
x4
1
b4
1
+
(1 + β2
1) + x2
1
4
b4
1
(3 + β2
1) + x2
2
4β2
1
b4
1
(1 + 3β2
Next, by using representation (4.38), we can write p1(x, t) in the form
1x4
β4
2
b4
1 (cid:19) .
1)(cid:19) .
(6.18)
(6.19)
(6.20)
(6.21)
(6.22)
(6.24)
(6.25)
4
b2
1
∆p0(x, t) = η(1)(t)(cid:18)−
p1(x, t) = 2η(1)(t)(cid:18)a40x4
−
Therefore, we obtain
1 + a22x2
1x2
2 + a04x4
1x2
2
1x4
a40β2
b2
1
1x4
2
1x2
a22β2
b2
1
−
−
a40x6
1
2 −
1x6
a04β2
2
b2
1 −
b2
1 (cid:19) .
1x2
a22x4
2
1 −
b2
1x4
2
a04x2
b2
1
(6.23)
∆p1(x, t) = 2η(1)(t)(cid:18)(12a40 + 2a22)x2
1 + (2a22 + 12a04)x2
2
12β2
1a40 + 12a22(1 + β2
1) + 12a04
x2
1x2
2
−
−
b2
1
a40(30 + 2β2
1) + 2a22
b2
1
23
x4
1 −
2a22β2
1 + a04(2 + 30β2
1)
b2
1
x4
2(cid:19).
We also use the following representations:
Ψj(x) = x2
1 + e2
j x2
2,
j = 1, 2.
Thus, applying the gradient operator, we simply get
and
It yields the following representation:
∇Ψ2(x) =(cid:0)2x1, 2e2
2x2(cid:1)
∇P0(x, t) =(cid:0)K−1∇p0(x,·)(cid:1) (t).
1 (cid:19)(cid:18)x2
1 (cid:19)(cid:19) (t) + 8x4
x2
1
1 −
b2
β2
1x2
2
b2
1
b2
1
b2
+
∇Ψ2(x) · ∇P0(x, t) = −8(cid:18)K−1(cid:20)η(1)(cid:18)1 −
= −8x2
1(cid:18)K−1(cid:18)η(1)
2x2
b2
1 (cid:19)(cid:19) (t) − 8e2
2(cid:18)K−1(cid:18)η(1)β2
2(cid:18)K−1(cid:18)η(1)β2
1 (cid:19)(cid:19) (t) + 8e2
1x2
1
1
b4
2θ2,2(t) + 8x4
+8(1 + e2
2)x2
=: −8x2
1θ2,0(t) − 8e2
2x2
(6.26)
2β2
1x2
e2
2
b2
1 (cid:19)(cid:21)(cid:19) (t)
1(cid:18)K−1(cid:18)η(1)
1 (cid:19)(cid:19) (t)
b4
1
1 (cid:19)(cid:19) (t)
2x4
2(cid:18)K−1(cid:18)η(1)β4
b4
1θ4,0(t) + 8(1 + e2
2)x2
1x2
2θ4,2(t) + 8e2
2x4
2θ4,4(t).
Here we have introduced the notation
Combining the above results we obtain the system of equations (4.42) -- (4.47).
θ2k,2l =(cid:0)K−1(cid:0)η(1)b−2k
1 β2l
1(cid:1)(cid:1) (t).
24
|
1503.04669 | 2 | 1503 | 2016-02-19T09:32:53 | Alzheimer's disease: a mathematical model for onset and progression | [
"physics.bio-ph",
"math.DS",
"physics.med-ph",
"q-bio.TO"
] | In this paper we propose a mathematical model for the onset and progression of Alzheimer's disease based on transport and diffusion equations. We regard brain neurons as a continuous medium, and structure them by their degree of malfunctioning. Two different mechanisms are assumed to be relevant for the temporal evolution of the disease: i) diffusion and agglomeration of soluble polymers of amyloid, produced by damaged neurons; ii) neuron-to-neuron prion-like transmission. We model these two processes by a system of Smoluchowski equations for the amyloid concentration, coupled to a kinetic-type transport equation for the distribution function of the degree of malfunctioning of neurons. The second equation contains an integral term describing the random onset of the disease as a jump process localised in particularly sensitive areas of the brain. Our numerical simulations are in good qualitative agreement with clinical images of the disease distribution in the brain which vary from early to advanced stages. | physics.bio-ph | physics |
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR
ONSET AND PROGRESSION
MICHIEL BERTSCH, BRUNO FRANCHI, NORINA MARCELLO, MARIA CARLA TESI,
AND ANDREA TOSIN
Abstract. In this paper we propose a mathematical model for the onset and
progression of Alzheimer's disease based on transport and diffusion equations.
We regard brain neurons as a continuous medium, and structure them by their
degree of malfunctioning. Two different mechanisms are assumed to be relev-
ant for the temporal evolution of the disease: i) diffusion and agglomeration
of soluble polymers of amyloid, produced by damaged neurons; ii) neuron-to-
neuron prion-like transmission. We model these two processes by a system of
Smoluchowski equations for the amyloid concentration, coupled to a kinetic-
type transport equation for the distribution function of the degree of malfunc-
tioning of neurons. The second equation contains an integral term describing
the random onset of the disease as a jump process localised in particularly
sensitive areas of the brain. Our numerical simulations are in good qualitative
agreement with clinical images of the disease distribution in the brain which
vary from early to advanced stages.
1. Introduction
Alzheimer's disease (AD) is one of the most common late life dementia, with huge
social and economic impact [6, 18, 24]. Its global prevalence, about 24 millions in
2011, is expected to double in 20 years [38]. In silico research, based on mathem-
atical modelling and computer simulations [1, 8, 10, 13, 17, 28, 36, 42], effectively
supplements in vivo and in vitro research. We present a multiscale model for the
onset and evolution of AD which accounts for the diffusion and agglomeration of
amyloid-β (Aβ) peptide (amyloid cascade hypothesis [16, 22]), and the spreading
of the disease through neuron-to-neuron transmission (prionoid hypothesis [7]).
Indeed, to cover such diverse facets of AD in a single model, different spatial and
temporal scales must be taken into account: microscopic spatial scales to describe
the role of the neurons, macroscopic spatial and short temporal (minutes, hours)
scales for the description of relevant diffusion processes in the brain, and large
temporal scales (years, decades) for the description of the global development of
AD. The way in which we combine distinct scales in a single model forms the core
and major novelty of the paper.
Following closely the biomedical literature on AD, we briefly describe the pro-
cesses which we shall include in our model. In the neurons and their interconnec-
tions several microscopic phenomena take place. It is largely accepted that beta
amyloid (Aβ), especially its highly toxic oligomeric isoforms Aβ40 and Aβ42, play
an important role in the process of the cerebral damage (the so-called amyloid cas-
cade hypothesis [22]). In this note we focus on the role of Aβ42 in its soluble form,
which recently has been suggested to be the principal cause of neuronal death and
eventually dementia [43]. At the level of the neuronal membrane, monomeric Aβ
peptides originate from the proteolytic cleavage of a transmembrane glycoprotein,
2010 Mathematics Subject Classification. 35M13, 35Q92, 92B99.
Key words and phrases. Alzheimer's disease, transport and diffusion equations, Smoluchowski
equations, numerical simulations.
1
2
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
the amyloid precursor protein (APP). By unknown and partially genetic reasons,
some neurons present an unbalance between produced and cleared Aβ (we refer to
such neurons as damaged neurons). In addition to this, it has been proposed that
neuronal damage spreads in the neuronal net through a neuron-to-neuron prion-like
propagation mechanism [7, 36].
On the other hand, macroscopic phenomena take place at the level of the cereb-
ral tissue. The monomeric Aβ produced by damaged neurons diffuses through the
microscopic tortuosity of the brain tissue and undergoes a process of agglomeration,
leading eventually to the formation of long, insoluble amyloid fibrils, which accu-
mulate in spherical deposits known as senile plaques. Moreover, soluble Aβ shows
a multiple neurotoxic effect:
it induces a general inflammation that activates the
microglia (the resident immune cells in the central nervous system) which in turn
secretes proinflammatory innate cytokines [15] and, at the same time, increases
intracellular calcium levels [13] yielding ultimately apoptosis and neuronal death.
The model we present is a conceptual interdisciplinary construction based on
clinical and experimental evidence, yielding in particular numerical simulations
and related graphs, that can be compared with time-dependent trajectories of AD
biomarkers (see e.g., [20, 21]). In particular, Figure 2 fits the core of the model
proposed in [20] for the temporal progression of the abnormalities in AD biomarkers,
which identifies two subsequent periods:
− a first period of β-amyloidosis characterised prevalently by reductions in
CSF Aβ42 and increased amyloid plaque formation (biomarkers of this first
period in our model correspond to CSF-Aβ42 and PIB-PET, Pittsburgh
compound B - Positron Emission Tomography);
− a second one characterised by neuronal dysfunction and neurodegeneration
(for this period we only take into account the structural MRI).
Of particular medical interest is the initial stage of the second period, which is
commonly referred to as Mild Cognitive Impairment (MCI): see e.g., [33].
2. Mathematical model
Highly toxic oligomeric isoforms of beta amyloid, Aβ40 and Aβ42, cause cereb-
ral damage. Here we restrict our attention to Aβ42 (shortly Aβ in the sequel)
in soluble form, generally considered the principal cause of neuronal death and
dementia [43]. Monomeric Aβ peptides originate from proteolytic cleavage of a
transmembrane glycoprotein, the amyloid precursor protein (APP). In AD, neur-
ons progressively present an unbalance between produced and cleared Aβ, but the
underlying mechanism is still largely unknown. On the other hand it was proposed
that neuronal damage spreads in the neural pathway through a neuron-to-neuron
prion-like propagation mechanism [7, 36].
Soluble Aβ diffuses through the microscopic tortuosity of the brain tissue and
undergoes an agglomeration process. Eventually this leads to the formation of
long, insoluble fibrils, accumulating in spherical deposits known as senile plaques.
Soluble Aβ has a multiple neurotoxic effect [13, 15]. In our model we do not enter
the details of the brain tissue, we neglect the action of the τ -protein, we simplify
the role of microglia and neglect its multifaceted mechanism (see e.g., [10] and [35]).
We simply assume that high levels of soluble amyloid are toxic for neurons.
We identify a portion of the cerebral tissue with a 3-dimensional region Ω and
x ∈ Ω indicates a generic point. Two temporal scales are needed to simulate the
longitudinal evolution of the disease over a period of years: a short (i.e., rapid) s-
scale (unit time coincides with hours) for the diffusion and agglomeration of Aβ [25],
and a long (i.e., slow) t-scale (unit time coincides with several months) for the
progression of AD, so ∆t = ε∆s for a small constant ε (cid:28) 1.
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION 3
We denote the molar concentration of soluble Aβ polymers of length m at point
x and time s by um(x, s), with 1 ≤ m < N . That of clusters of oligomers of length
≥ N (fibrils) is denoted by uN (x, s) and may be thought as a medical parameter
(the plaques), clinically observable through PIB-PET ( [30]).
To model the aggregation of Aβ m-polymers (1 < m < N ) we follow [1],
[variation in (short) time] = [diffusion] + [agglomeration],
which, in mathematical terms, leads to the Smoluchowski equation with diffusion:
1
2
m−1(cid:88)
N(cid:88)
.
(1)
∂sum = dm∇2um +
aj,m−jujum−j − um
am,juj
j=1
j=1
where dm > 0, m = 1, . . . , N , and ai,j = aj,i > 0, i, j = 1, . . . , N (giving the
factor 1
2 in (1)).
We refer to [1, 12] for an extensive discussion of (1). For reasons related to
the model, we can assume that the diffusion coefficients dm are small when m is
large, since big assemblies do not move. In fact, the diffusion coefficient of a soluble
peptide scales approximately as a reciprocal of the cube root of its molecular weight
(see [14] and also [29]).
Applications of the Smoluchowski equation to the description of the agglomera-
tion of Aβ amyloid appear in [28]. In this paper, the authors compare experimental
data, obtained in vitro, with numerical simulations based on the Smoluchowski
equation (without diffusion) in order to describe the process leading to insoluble
fibril aggregates from soluble amyloid. The form of the coefficients ai,j (the coagu-
lation rates) we use has been considered by in vitro Murphy & Pallitto (see [28]
and [32]). According to formula (13) in [28], the coagulation rates in silico in our
equations take the form
(2)
ai,j = const.
1
i + j
·
+
ln(j/d) + νj
j
i
(cid:18) ln(i/d) + νi
(cid:19)
,
where i, j are the lengths of the fibrils, d is their diameter, and νi = 0.312 +
0.565(i/d)−1 − 0.1(i/d)−2.
The physical arguments leading to formula (2) rely on sophisticated statistical
mechanics considerations (see also [40]).
νi = ν for = 1, . . . , N − 1. Thus we can replace (2) by
Since d can be assumed very small, without loss of generality we can assume
(cid:19)
·(cid:0)ν + ln d + O(ln N )(cid:1).
(3) ai,j = const.
1
i + j
·
+
ln(j/d) + ν
j
=
1
ij
i
(cid:18) ln(i/d) + ν
Since N is finite, in our numerical simulations we use a slightly approximate form
of these coefficients, taking
(4)
ai,j = α
1
ij
, where α > 0.
Smoluchowski equations with diffusion have already been considered in the lit-
erature (without reference to Aβ amyloid and Alzheimer's disease) with diverse
boundary conditions: see for instance [9] for a general introduction, and [23, 2, 44,
4, 3].
term F:
Neurons produce Aβ monomers, whence the equation for u1 contains a source
(5)
∂su1 = d1∇2u1 − u1
a1,juj + F.
N(cid:88)
j=1
4
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
Since fibrils are assumed not to move, the equation for uN has no diffusion term,
and takes the form (see (4) in [1]):
(6)
∂suN =
1
2
aj,kujuk.
(cid:88)
j+k≥N
k, j<N
(cid:88)
j+k≥N
N(cid:88)
j=1
N−1(cid:88)
m−1(cid:88)
j=1
m−1(cid:88)
(cid:88)
It is coherent with experimental data to assume aN,N = 0 for large N . This is
equivalent to saying that large oligomers do not aggregate with each other.
The justification of the condition j, k < N in (6) requires a few more words. In
fact, we must remember that the meaning of uN differs from that of um, m < N ,
as well as the identity
(7)
1
2
j+k≥N,k<N,j<N
aj,kujuk =
1
2
aj,kujuk − uN
aN,juj.
The idea is that uN should describe the sum of the densities of all the "large" assem-
blies. We assume that large assemblies exhibit all the same coagulation properties
and do not coagulate with each other. Let us briefly show how (6) is obtained: we
start by writing the exact Smoluchowski equation for all m ≥ 1 using um instead
of um in order to avoid confusion, i.e. nothing but the PDE in (1) with m ranging
from 2 to ∞. We have
N(cid:88)
j=1
(8)
um = dm∇2 um − um
∂
∂t
am,j uj +
1
2
aj,m−j uj um−j,
where, coherently with our assumptions, we assume
i) dm = dN for m ≥ N ;
ii) am,j = aN,j for m ≥ N . In particular, if m, j ≥ N , am,j = aN,j = aN,N =
0.
Therefore, if m ≥ N , (8) becomes
∂
∂t
um = dN∇2 um − um
(9)
aN,j uj +
Now we sum up (9) for m ≥ N , and we set for a while v :=(cid:80)
m−1(cid:88)
m≥N um. We want
to show precisely that v satisfies the equation (6) (satisfied by uN ). By i), we have
aj,m−j uj um−j,
N−1(cid:88)
(cid:88)
j=1
j=1
1
2
um
aN,j uj +
ai,m−i ui um−i
1
2
m≥N
i=1
∂v
∂t
= dN∇2v − (cid:88)
m≥N
:= dN∇2v − I1 +
j=1
1
2
I2.
It is clear that
I1 =
(cid:88)
m≥N
N−1(cid:88)
j=1
N−1(cid:88)
j=1
aN,j uj,
aN,j uj = v
um
that is precisely the second term in (7), since aN,N = 0. As for I2, if we set j := i
and k := m − i, we obtain the first term in (7). Finally, if set um = um for m < N
and uN = v we recover the PDE in (6), as desired.
We model the degree of malfunctioning of a neuron with a parameter a ranging
from 0 to 1: a close to 0 stands for "the neuron is healthy" whereas a close to
1 for "the neuron is dead". This parameter, although introduced for the sake of
mathematical modelling (see also [36]), can be compared with medical images from
Fluorodeoxyglucose PET (FDG-PET [27]).
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION 5
Given x ∈ Ω, t ≥ 0, and a ∈ [0, 1],
f (x, a, t) da
indicates the fraction of neurons close to x with degree of malfunctioning at time t
between a and a + da. The progression of AD occurs at the long time scale t, over
decades, and is determined by the deterioration rate, v = v(x, a, t), of the health
state of the neurons:
(10)
∂tf + ∂a(f v[f ]) = 0.
Here v[f ] indicates that the deterioration rate depends on f itself. The onset of
AD will be included in a subsequent step.
We assume that
(cid:90)(cid:90)
(11)
v[f ] =
Ω×[0, 1]
K(x, a, y, b)f (y, b, t) dy db + S(x, a, u1(x, s), . . . , uN−1(x, s)).
The integral term describes the possible prion-like propagation of AD through the
neural pathway. Malfunctioning neighbours are harmful for a neuron's health state,
while healthy ones are not:
K(x, a, y, b) ≥ 0
K(x, a, y, b) = 0
∀ x, y ∈ Ω, a, b ∈ [0, 1],
if a > b.
Typically
with, for example,
K(x, a, y, b) = G(x, a, b)H(x, y)
G(x, a, b) = CG(b − a)+,
H(x, y) = h(x − y),
(cid:82)
where (·)+ denotes the positive part (x+ := max{0, x}) while h(r) is a non-
negative and decreasing function, which vanishes at some r = r0 and satisfies
y<r0
h(y) dy = 1. In the limit r0 → 0, (11) reduces to
(12)
v[f ] =
G(x, a, b)f (x, b, t) db + S(x, a, u1(x, s), . . . , uN−1(x, s)).
(cid:90) 1
0
Since we aim at a minimal effective model, we avoid precise assumptions on the
underlying biological processes expressed by K.
The term S ≥ 0 in (11) and (12) models the action of toxic Aβ oligomers,
ultimately leading to apoptosis. For example
(13)
S = CS (1 − a)
mum(x, s) − U
m=1
The threshold U > 0 indicates the minimal amount of toxic Aβ needed to damage
neurons, assuming that the toxicity of soluble Aβ-polymers does not depend on m.
In reality length dependence has been observed [31], but, to our best knowledge,
quantitive data are only available for very short molecules (see [31, Table 2]). For
long molecules any analytic expression would be arbitrary.
Since Aβ monomers are produced by neurons and the production increases if
neurons are damaged, we choose in (5)
(14)
F = F[f ] = CF
(µ0 + a)(1 − a)f (x, a, t) da.
The small constant µ0 > 0 accounts for Aβ production by healthy neurons (dead
neurons do not produce amyloid).
0
(cid:90) 1
(cid:32)N−1(cid:88)
(cid:33)+
6
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
To describe the onset of AD we assume that in small, randomly chosen parts
of the cerebral tissue, concentrated for instance in the hippocampus, the degree of
malfunctioning of neurons randomly jumps to higher values due to external agents
or genetic factors. This leads to an additional term in the equation for f ,
where
(cid:18)(cid:90) 1
∂tf + ∂a (f v[f ]) = J[f ],
(cid:19)
J[f ] = η
P (t, a∗ → a)f (x, a∗, t) da∗ − f (x, a, t)
(15)
P (t, a∗ → a) is the probability to jump from state a∗ to state a ∈ [0, 1] (obviously,
P (t, a∗ → a) = 0 if a < a∗), χ(x, t) describes the random jump distribution, and η
is the jump frequency. In most of our numerical tests we choose
χ(x, t).
0
P (t, a∗ → a) ≡ P (a∗ → a) =
2
1 − a∗
0
if a∗ ≤ a ≤ 1+a∗
otherwise,
2
i.e., we neglect randomness and we set χ(x, t) ≡ χ(x) concentrated in the hippo-
campus. For a simulation with a random jump distribution, see Figure 10.
To model the phagocytic activity of the microglia as well as other bulk clearance
processes [19], we add to (1) and (5) a term −σmum, where σm > 0. This leads to
the system
∂tf + ∂a (f v[f ]) = J[f ]
ε∂tu1 = d1∇2u1 − u1
ε∂tum = dm∇2um +
ε∂tuN =
1
2
(cid:80)
j+k≥N
k, j<N
N(cid:80)
m−1(cid:80)
N(cid:80)
j=1
j=1
1
2
− um
j=1
aj,kujuk,
(16)
a1,juj + F[f ] − σ1u1
aj,m−jujum−j
am,juj − σmum
(2 ≤ m < N )
where v[f ] is given by (11) or (12) (with s replaced by ε−1t), F[f ] by (14), and J[f ]
by (15). Since we are interested in longitudinal modelling, we assume that initially,
at t = 0, the brain is healthy, with a small uniform distribution of soluble amyloid.
3. Problem setting and discretisation of the equations
In this section we detail the structure of the domain and the boundary conditions
which we will use, in the next Section 4, to produce numerical simulations. We also
discuss the discretisation of the equations (16).
3.1. Physical domain and boundary conditions. We consider the two-dimen-
sional transverse section of the brain illustrated in Fig. 1. Since approximating a real
brain section is a quite complicated issue, for the sake of simplicity we schematise it
as a box Ω ⊂ R2, Ω = [0, Lx]×[0, Ly], with two inner rectangular holes representing
the sections of the cerebral ventricles. We also identify, close to the front part of the
ventricles, the two sections of the hippocampus, which we represent as two small
circles. Unlike the cerebral ventricles, the sections of the hippocampus are meant
as actual portions of the domain Ω, not as holes.
On the outer boundary of Ω, say ∂Ωout, we assume vanishing normal polymer
flow. Therefore we impose a homogeneous Neumann condition for the diffusing
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION 7
Figure 1. Left: A real transverse section of the brain (repro-
duced from [26] with kind permission of the publisher). Right:
Two-dimensional schematisation for numerical purposes. Black
dots are the internal nodes of the numerical grid, where discretised
equations are solved, while white dots are boundary nodes, where
boundary conditions are imposed.
amyloid polymers:
(17)
− dm
∇um · n = 0
on ∂Ωout, m = 1, . . . , N − 1,
n being the outward normal unit vector to ∂Ωout. Notice that no boundary condi-
tion is required for the concentration uN of the fibrillar amyloid, since its equation
does not feature space dynamics (cf. the last equation in (16)).
On the inner boundary of Ω, say ∂Ωin, that is the boundaries of the cerebral
ventricles, we model the removal of Aβ from cerebrospinal fluid (CSF) through
the choroid plexus (cf. [19, 39]) by an outward polymer flow proportional to the
concentration of the amyloid. For this, we impose a Robin boundary condition of
the form:
(18)
− dm
∇um · n = βum on ∂Ωin, m = 1, . . . , N − 1,
with β > 0 a constant.
We discretise Ω by means of a two-dimensional structured orthogonal grid, whose
points have coordinates xi,j = (xi, yj) = (i∆x, j∆y) with ∆x = Lx/Nx, ∆y =
Ly/Ny, Nx, Ny being the numbers of discretisation points in the x and y-direction,
respectively, and i = 0, . . . , Nx, j = 0, . . . , Ny. See Fig. 1. We also introduce a
time lattice tn = n∆t, n = 0, 1, 2, . . . .
i,j ≈ um(xi,j, tn) denote an approximation of the concentration of
the m-polymers of amyloid in the point xi,j ∈ Ω at time tn, on the numerical grid
Letting (um)n
8
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
the Neumann boundary condition (17) becomes simply:
(um)n
(um)n
Nx,j = (um)n
(um)n
0,j = (um)n
1,j
Nx−1,j
i,0 = (um)n
i,1
= (um)n
i,Ny−1
(um)n
i,Ny
j = 1, . . . , Ny − 1
i = 1, . . . , Nx − 1
(cid:27)
(cid:27)
m = 1, . . . , N − 1.
Concerning the Robin boundary condition (18), we discretise the components of
the gradient via the forward Euler formula, then we take into account the orientation
of the vector n as indicated in Fig. 1 to find:
− along the left boundary of each cerebral ventricle
where ib ∈ {0, . . . , Nx} denotes the grid index in the x-direction such that
xib = ib∆x is the abscissa of the boundary;
− along the right boundary of each cerebral ventricle
− along the lower boundary of each cerebral ventricle
where jb ∈ {0, . . . , Ny} denotes the grid index in the y-direction such that
yjb = jb∆y is the ordinate of the boundary;
− along the upper boundary of each cerebral ventricle
(um)n
ib,j =
(um)n
ib−1,j
1 + β∆x/dm
(um)n
ib,j =
(um)n
ib+1,j
1 + β∆x/dm
(um)n
i,jb
=
(um)n
i,jb−1
1 + β∆y/dm
,
;
,
(um)n
i,jb
=
(um)n
i,jb+1
1 + β∆y/dm
.
3.2. Discretisation of the Smoluchowski equations. In order to approximate
the equations for the um's, m = 1, . . . , N − 1, we use a fractional step procedure in
time: first we solve the diffusion and reaction parts, then we add the coagulation
and possibly the source (for u1) parts.
Adopting a Finite Difference discretisation of the Laplace operator ∇2 we obtain
the scheme:
(um)∗
i,j = (um)n
i,j + ∆t
dm
(um)n
i−1,j−2(um)n
∆x2
i,j +(um)n
i+1,j
(cid:16)
(cid:17)
− σm(um)n
i,j
(cid:18)
(um)n+1
i,j = (um)∗
i,j + ∆t
+dm
m−1(cid:80)
1
2
h=1
(um)n
i,j−1−2(um)n
∆y2
i,j +(um)n
i,j+1
ah,m−h(uh)∗
i,j(um−h)∗
i,j
(cid:19)
−(um)∗
i,j
am,h(uh)∗
i,j
,
N(cid:80)
h=1
where (um)∗
i,j denotes the temporary solution computed after the first fractional
time step. For an alternative Finite Element discretisation of Smoluchowski equa-
tions see e.g., [1].
The scheme above applies to all inner nodes xi,j of the numerical grid (that means
1 ≤ i ≤ Nx − 1, 1 ≤ j ≤ Ny − 1 excluding furthermore the indexes ib, jb identifying
the inner boundary ∂Ωin) and to the Aβ m-polymers with m = 2, . . . , N − 1.
Because of the adopted approximation of the diffusion part, the following constraint
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION 9
on the time and space steps has to be enforced for the stability of the numerical
scheme:
(19)
∆t ≤
4
· min{∆x2, ∆y2}
.
max
1≤m≤N−1
dm
For m = 1 the equation is discretised in a similar way but for the addition of
the source term F. We refer the reader to the next subsection for discretisation
methods of the integral contained in it.
Finally, for m = N the equation is actually an ODE, which we approximate by
the explicit Euler formula:
(uN )n+1
i,j = (uN )n
i,j +
∆t
2
ah,k(uh)n
i,j(uk)n
i,j.
(cid:88)
h+k≥N
h, k<N
3.3. Discretisation of the equation for f . In the interval [0, 1], which consti-
tutes the domain of the variable a, we introduce a Finite Volume partition made
of Na cells of the form [ak−1/2, ak+1/2] with central point ak =(cid:0)k − 1
(cid:1) ∆a, where
∆a = 1
Na
in (16) using again a fractional step procedure in time.
. The cell index k runs from 1 to Na. Then we discretise the first equation
2
(20)
(cid:0)f n
i,j,k ≈ f (xi,j, ak, tn), we have:
First, we solve the homogeneous transport part by means of the push-forward
scheme introduced by e.g., [34, 41], which is particularly suited to deal with non-
local fluxes. Denoting f n
i,j,k − ∆t
∆a
f∗
i,j,k = f n
i,j,k ≈ v(xi,j, ak, tn) indicates an approximation of the deterioration rate of
where vn
the neurons and (·)− is the negative part (x− := max{0, −x}). Here we compute
vn
i,j,k by approximating the integral contained in the expression (12) via a zeroth
order Euler formula and then adding the expression (13):
i,j,k+1)−(cid:1) ,
i,j,k−1)+ − f n
i,j,k − f n
i,j,kvn
i,j,k−1(vn
i,j,k+1(vn
vn
i,j,k =
G(xi,j, ak, ah)f n
i,j,h∆a + CS (1 − ak)
m(um)n
i,j − U
.
If the form G(x, a, b) = CG(b − a)+ is used then in the formula above we simply
have G(xi,j, ak, ah) = CG(ah − ak)+.
The stability of the scheme (20) requires that the grid steps ∆a, ∆t be linked
(cid:32)N−1(cid:88)
m=1
(cid:33)+
Na(cid:88)
h=1
by the following CFL condition:
(21)
∆t ≤
∆a
vn
i,j,k .
max
i, j, k
(cid:32) Na(cid:88)
Second, we update the values f∗
i,j,k by including the jump process:
i,j,k = f∗
f n+1
h f∗
P k
i,j,h∆a − f∗
i,j,k
χn
i,j,
i,j,k + η∆t
h := P (ah → ak) and χn
h=1
where we have denoted P k
i, j := χ(xi,j, tn).
(cid:33)
3.4. Final choice of the time step. On the whole, the time step of the com-
plete numerical scheme has to comply with both the parabolic and the hyperbolic
constraints (19), (21), respectively. Therefore, it is ultimately chosen as:
4
∆t ≤ min
· min{∆x2, ∆y2}
max
1≤m≤N−1
dm
,
∆a
vn
i,j,k
max
i, j, k
10
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
at each time iteration of the numerical scheme.
3.5. Computing physiological indicators. Several macroscopic (aggregate) quant-
ities can be computed out of the results of model (16). In the next Section 4 the
outputs of the simulations will be discussed in terms of a few of such quantities,
which can be directly compared with real clinical images and known qualitative
time evolution of Alzheimer's disease.
The macroscopic distribution of neuron malfunctioning A = A(x, t) is computed
over the cerebral domain Ω as the local average of the degree of malfunctioning a:
A(x, t) :=
af (x, a, t) da,
which is numerically approximated as
0
A(xi,j, tn) ≈ An
i,j =
Na(cid:88)
k=1
akf n
i,j,k.
Following [21], we relate then the local brain atrophy φ(x, t) to the average neuron
malfunctioning A as:
(cid:90) 1
(cid:26)
(cid:26)
(cid:27)
,
A(x, t) − A0
1 − A0
(cid:27)
.
i,j − A0
An
1 − A0
φ(x, t) := max
0,
φn
i,j = max
0,
A0 ∈ (0, 1) being a threshold of malfunctioning over which the brain is considered
locally atrophic. The corresponding numerical approximation is
Next we define the global brain atrophy in time Φ = Φ(t) as the average of φ over
the whole domain Ω, i.e.,
Ω denoting the area of Ω, which is numerically approximated as
Φ(t) :=
φ(x, t) dx,
1
Ω
(cid:90)
Ny−1(cid:88)
Nx−1(cid:88)
Ω
Φn :=
1
Ω
φn
i,j∆x∆y,
j=0
In this formula we conventionally consider φn
belong to the domain Ω, i.e., if it is a point inside the cerebral ventricles.
i,j = 0 if the grid point xi,j does not
i=0
The total concentration of soluble amyloid US = US(t) in the brain occipital
region, to be related to the Aβ concentration found in the cerebrospinal fluid by
clinical exams (CSF Aβ), is given by:
(cid:90)
N−1(cid:88)
Ω
m=1
US(t) :=
1
Ω
mum(x, t) dx,
where Ω ⊂ Ω is a subdomain located in the bottom part of Ω, entirely contained
in the region below the cerebral ventricles. Assuming for simplicity that it is a
rectangle as well, whose grid coordinates are comprised between two indexes 0 <
i1 < i2 < Nx in the x-direction and between j = 0 and j = j > 0 in the y-direction,
we obtain the numerical values of US as:
i2−1(cid:88)
j−1(cid:88)
N−1(cid:88)
i=i1
j=0
m=1
U n
S :=
1
Ω
m(um)n
i,j∆x∆y =
1
(i2 − i1)j
i2−1(cid:88)
j−1(cid:88)
N−1(cid:88)
i=i1
j=0
m=1
m(um)n
i,j,
where we have used that Ω = (i2 − i1)j∆x∆y.
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION11
Finally, the average quantity of brain Aβ deposits in time is:
UN (t) :=
which is naturally discretised as
N uN (x, t) dx,
(cid:90)
Nx−1(cid:88)
Ω
i=0
Ny−1(cid:88)
1
Ω
1
Ω
by letting conventionally (uN )n
UN (tn) ≈ U n
N (uN )n
j=0
i,j∆x∆y
N =
i,j = 0 if xi,j (cid:54)∈ Ω (inside the cerebral ventricles).
4. Numerical results
To begin with, we provide a typical output of the numerical simulations.
In
Figure 2 we plot the evolution of three crucial biomarkers of AD (as a function of
the computational time):
− the CSF Aβ42 (purple dashed curve);
− the average quantity of brain Aβ42 deposits (red solid curve);
− the global brain atrophy (blue dash-dot curve).
All curves are normalised to their maxima. The values of the constants used in the
simulation are specified in the figure caption.
Figure 2. Graph for the following constants: β = 1, D = 0.01,
α = 10, ε = 0.1, T = 100, N = 50, CG = 0.1, CS = 0.001,
CF = 10, r0 = 0.0, U = 0.1, µ0 = 0.01, η = 1, and σm = 1/m.
The level of Aβ42-deposition (red solid curve) grows rapidly, reaches its max-
imum, and then stabilises. The purple dashed curve, corresponding to CSF-Aβ42,
decreases after having reached a peak. The blue dash-dot curve corresponds to the
brain atrophy and increases in time as expected. The graphs in Figure 2 can be
well illustrated by the following quote from [20]:
The initiating event in AD is related to abnormal processing of
β-amyloid peptide, ultimately leading to formation of Aβ plaques
in the brain. This process occurs while individuals are still cog-
nitively normal. Biomarkers of brain β-amyloidosis are reductions
in CSF Aβ42 and increased amyloid PET tracer retention. After
a lag period, which varies from patient to patient, neuronal dys-
function and neurodegeneration become the dominant pathological
0 0.2 0.4 0.6 0.8 1 0 20 40 60 80 100 120 140Normalized valuesComputational timeCSF Aβ42Brain Aβ42 depositsGlobal brain atrophy12
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
processes. Biomarkers of neuronal injury and neurodegeneration
are increased CSF tau and structural MRI measures of cerebral
atrophy. Neurodegeneration is accompanied by synaptic dysfunc-
tion, which is indicated by decreased fluorodeoxyglucose uptake
on PET. We propose a model that relates disease stage to AD
biomarkers in which Aβ biomarkers become abnormal first, be-
fore neurodegenerative biomarkers and cognitive symptoms, and
neurodegenerative biomarkers become abnormal later, and correl-
ate with clinical symptom severity.
The plots we obtain should be compared with the clinical graphs in [21],
[45],
[11] and [37]. For the reader's convenience we reproduce
[5], and with the data of
here a picture from [21], see Figure 3, and a picture from [45], see Figure 4.
Figure 3. Fig. 6 reproduced from [21] with kind permission of the publisher.
Figure 4. Fig. 4 reproduced from [45] with kind permission of the publisher.
There is a satisfactory agreement between the plots of the qualitative temporal
behaviour of the biomarkers and those obtained from clinical data. Observe that
not only the shapes of the curves are comparable (CSF Aβ corresponds to CSF
Aβ42, Brain Aβ deposits correspond to Amyloid PET and Global brain atrophy
corresponds to MRI + FDG PET)), but also the temporal order of the events is in
good agreement with clinical data.
Obviously the details of the numerical output depend on the choice of the con-
stants used in the mathematical model. Performing a considerable amount of nu-
merical runs with different values of the constants in the model, we have reached
the conclusion that, at least qualitatively, the behaviour of the solutions does not
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION13
depend on the precise choice of those constants, as long as their variation is re-
stricted to reasonable ranges. In other words, the values of the constants taken in
Figure 2 can be considered as an indication for the order of magnitude. For ex-
ample, the longitudinal graphs of the biomarkers CFS-Aβ, brain Aβ deposits and
brain atrophy are - in this sense - qualitatively stable under variations of CS , CF ,
CG and α.
It is particularly instructive to consider the constants U in (13) and β in (18).
We recall that U is a threshold value for the minimal amount of toxic Aβ necessary
to damage neurons (see (13)). In Figure 2 we have used the value U = 0.1, but
if we make it considerably larger, for example U = 1 (the remaining constants are
unchanged), then the threshold becomes so high that the illness does not develop
at all.
The constant β enters the model through condition (18) at the boundary of the
cerebral ventricles. Smaller values of β mean that less Aβ is removed from the
CSF through the choroid plexus. Figure 5 shows what happens if we change it
into β = 0.01: the three curves are moved to the left and become steeper: the
illness starts earlier and develops faster. Recalling that in Figures 2 and 5 we have
plotted values which are normalised with respect to their maximal values, one could
wonder how the latter ones depend on β. It turns out that the maximal values of
CSF Aβ and the brain atrophy are essentially independent of β. The Aβ deposits
(the plaques) however increase by a factor 6 if β is changed from 1 to 0.01. This
result is compatible with our modelling Ansatz (in accordance with the medical
literature) that plaques are not toxic (even healthy brains may contain plaques).
Figure 5. Graph for the following constants: β = 0.01, D = 0.01,
α = 100, ε = 0.1, T = 100, N = 50, CG = 0.1, CS = 0.001,
CF = 10, r0 = 0.0, U = 0.1,µ0 = 0.01, η = 1, and σm = 1/m.
The comparison of the cases β = 1 and β = 0.01 becomes even clearer when
we create spatial plots of f and of the distribution and density of the cerebral
plaques at fixed computational times t = T . The plots of f at different times are
meant to be compared with FDG-PET images (see e.g., [11]). More precisely, we
take a schematic image of a transverse section of the brain and attribute different
colors (varying from red to blue) to those parts of the brain where probabilistic-
ally the level of malfunctioning lies in different ranges. As in the FDG-PET, the
0 0.2 0.4 0.6 0.8 1 0 20 40 60 80 100 120 140Normalized valuesComputational timeCSF Aβ42Brain Aβ42 depositsGlobal brain atrophy14
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
red corresponds to a healthy tissue. Here AD originates only from the hippocam-
pus and propagates, at the beginning, along privileged directions (such as those
corresponding to denser neural bundles) mimicked by two triangles.
In Figures 7 and 8 we compare plots of f at, respectively, times T = 34 and
T = 52 for the two different values of β = 0.01 and β = 1. Figures 7 and 8 do not
only confirm the temporal acceleration of the development of the illness for smaller
values of β, but also show that the spatial pattern and heterogeneity become less
evident as β becomes smaller. Since experimental data suggest a strong spatial
heterogeneity of the illness, this could indicate the potential importance of the
removal of Aβ through the choroid plexus to slow down the temporal development
of AD.
In Figure 9 we plot the plaques' distribution for the two different values of
β = 0.01 and β = 1 and at T = 52. This figure confirms the strong increase of the
plaques when β becomes smaller.
We stress that, though our images represent a mean value of brain activity
instead of a single patient's brain activity, still they show a good agreement with
clinical neuroimaging: compare Figures 7 and 8 with Figure 6 below.
Figure 6. FDG -PET images showing patterns of metabolic
activity: an elderly individual with no dementia (left) and with
AD (right). Reproduced from [26] with permission.
Figure 7. Neuron malfunctioning: β = 0.01 (left), β = 1 (right),
T = 34.
Looking for more realistic images, we have to take into account randomness of the
spatial distribution of the sources of the disease. For example, we have performed
0 0.8 0 1 0 0.2 0.4 0.6 0.8 1 0 0.8 0 1 0 0.2 0.4 0.6 0.8 1ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION15
Figure 8. Neuron malfunctioning: β = 0.01 (left), β = 1 (right),
T = 52.
Figure 9. Density of plaques for β = 0.01 (left), β = 1 (right),
T = 52.
some runs where the AD does not only originate from the hippocampus, but also
from several sources of Aβ randomly distributed in the occipital part of the brain.
We report the outputs of such runs in Figure 10. The random distributed sources
appear as the small blue spots.
5. Discussion and future research directions
We have presented a new mathematical model for the onset and evolution of AD.
The model is characterised by a high level of flexibility, which potentially allows one
to simulate different modelling hypotheses and compare them with clinical data. In
fact, the model provides a flexible tool to test in the future alternative hypotheses
on the evolution of the disease. In the paper we have chosen some specific aspects
of the illness, such as the aggregation, diffusion and removal of Aβ, the possible
spread of neuronal damage in the neural pathway, and, to describe the onset of AD,
a random neural deterioration mechanism. Numerical simulations are compared
with clinical data and, although oversimplified and restricted to a 2-dimensional
rectangular section of the brain, they are in good qualitative agreement with the
0 0.8 0 1 0 0.2 0.4 0.6 0.8 1 0 0.8 0 1 0 0.2 0.4 0.6 0.8 1 0 0.8 0 1 0 0.5 1 1.5 2 2.5 0 0.8 0 1 0 0.5 1 1.5 2 2.516
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
Figure 10. Neuron disease with random sources with β = 1 at
T = 47
spread of the illness in the brain at various stages of its evolution. In particular,
our model captures the cerebral damage in the early stage of MCI.
There are multiple future research developments in quite different directions,
each of which requires substantial research efforts. We mention some of them.
Further development of the model is needed and should be carefully guided by
clinical data. The constants appearing in the equations should be well calibrated
to optimise quantitative agreement with clinical data. Simulations should become
more realistic, in a three-dimensional domain which matches the geometric charac-
teristics of the brain.
The true challenge in AD research is a breakthrough which allows one to develop
effective therapies to stop or slow down the evolution of AD, possibly in an early
stage of the illness. Also effective mathematical models can give a contribution
in this direction. For example, a certain sensibility of the numerical output to
the value of the constant β in (18), which models the removal of Aβ through the
choroid plexus, spontaneously leads to the question whether dialysis-mechanisms
can be introduced to enhance Aβ-removal artificially. Most probably, a serious
answer to this question requires, in addition to a detailed comparison with clinical
data, a more refined modelling of the removal which takes into account the transport
of soluble Aβ by the cerebral fluid.
Finally, some mathematical effort is necessary to check the mathematical cor-
rectness (well-posedness) of the model.
Acknowledgments
B. F. and M. C. T. are supported by a grant of the University of Bologna (Ricerca
Fondamentale Orientata). B. F. and M. C. T. are supported by GNAMPA (Gruppo
Nazionale per l'Analisi Matematica, la Probabilit`a e le loro Applicazioni) of INdAM
(Istituto Nazionale di Alta Matematica), Italy. A. T. acknowledges that this work
has been written within the activities of GNFM (Gruppo Nazionale per la Fisica
Matematica) of INdAM (Istituto Nazionale di Alta Matematica), Italy.
0 0.8 0 1 0 0.2 0.4 0.6 0.8 1ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION17
References
1. Y. Achdou, B. Franchi, N. Marcello, and M. C. Tesi, A qualitative model for aggregation and
diffusion of β-amyloid in Alzheimer's disease, J. Math. Biol. 67 (2013), no. 6-7, 1369–1392.
MR 3125569
2. H. Amann, Coagulation-fragmentation processes, Arch. Ration. Mech. Anal. 151 (2000), no. 4,
339–366.
3. H. Amann and C. Walker, Local and global strong solutions to continuous coagulation-
fragmentation equations with diffusion, J. Differential Equations 218 (2005), no. 1, 159–186.
4. H. Amann and F. Weber, On a quasilinear coagulation-fragmentation model with diffusion,
Adv. Math. Sci. Appl. 11 (2001), no. 1, 227–263.
5. R. J. Bateman, C. Xiong, T. L. S. Benzinger, A. M. Fagan, A. Goate, N. C. Fox, D. S. Marcus,
N. J. Cairns, X. Xie, T. M. Blazey, D. M. Holtzman, A. Santacruz, V. Buckles, A. Oliver,
K. Moulder, P. S. Aisen, B. Ghetti, W. E. Klunk, E. McDade, R. N. Martins, C. L. Masters,
R. Mayeux, J. M. Ringman, M. N. Rossor, P. R. Schofield, R. A. Sperling, S. Salloway, and
J. C. Morris, Clinical and biomarker changes in dominantly inherited Alzheimer's disease,
New Engl. J. Med. 367 (2012), no. 9, 795–804.
6. K. Blennow, M. J. de Leon, and H. Zetterberg, Alzheimer's disease, Lancet Neurol. 368
(2006), 387–403.
7. Heiko Braak and Kelly Del Tredici, Alzheimer's pathogenesis:
is there neuron-to-neuron
propagation?, Acta Neuropathol. 121 (2011), no. 5, 589–595 (English).
8. L. Cruz, B. Urbanc, S. V. Buldyrev, R. Christie, T. G´omez-Isla, S. Havlin, M. McNamara,
H. E. Stanley, and B. T. Hyman, Aggregation and disaggregation of senile plaques in Alzheimer
disease, P. Natl. Acad. Sci. USA 94 (1997), no. 14, 7612–7616.
9. R. L. Drake, A general mathematical survey of the coagulation equation, Topics in Current
Aerosol Research (Part 2), International Reviews in Aerosol Physics and Chemistry (G. M.
Hidy and J. R. Brock, eds.), Pergamon Press, Oxford, UK, 1972, pp. 203–376.
10. Leah Edelstein-Keshet and Athan Spiross, Exploring the formation of Alzheimer's disease
senile plaques in silico, J. Theor. Biol. 216 (2002), no. 3, 301–326.
11. A. S. Fleisher, K. Chen, Y. T. Quiroz, L. J. Jakimovich, M. G. Gomez, C. M. Langois, J. B. S.
Langbaum, N. Ayutyanont, A. Roontiva, P. Thiyyagura, W. Lee, H. Mo, L. Lopez, S. Moreno,
N. Acosta-Baena, M. Giraldo, G. Garcia, R. A. Reiman, M. J. Huentelman, K. S. Kosik, P. N.
Tariot, F. Lopera, and E. M. Reiman, Florbetapir PET analysis of amyloid-β deposition in
the presenilin 1 E280A autosomal dominant Alzheimer's disease kindred: a cross-sectional
study, Lancet Neurol. 11 (2012), no. 12, 1057–1065.
12. B. Franchi and M. C. Tesi, A qualitative model for aggregation-fragmenattion and diffusion of
β-amyloid in Alzheimer's disease, Rend. Semin. Mat. Univ. Politec. Torino 70 (2012), no. 1,
75–84.
13. T. A. Good and R. M. Murphy, Effect of β-amyloid block of the fast-inactivating K+ channel
on intracellular Ca2+ and excitability in a modeled neuron, P. Natl. Acad. Sci. USA 93 (1996),
15130–15135.
14. Geoffrey J. Goodhill, Diffusion in axon guidance, European Journal of Neuroscience 9 (1997),
no. 7, 1414 – 1421.
15. W. S. T. Griffin, J. G. Sheng, M. C. Royston, S. M. Gentleman, J. E. McKenzie, D. I.
Graham, G. W. Roberts, and R. E. Mrak, Glial-neuronal interactions in Alzheimer's disease:
the potential role of a cytokine cycle in disease progression, Brain Pathol. 8 (1998), no. 1,
65–72.
16. C. Haass and D. J. Selkoe, Soluble protein oligomers in neurodegeneration: lessons from the
Alzheimer's amyloid beta-peptide, Nat. Rev. Mol. Cell. Biol. 8 (2007), no. 2, 101–112 (eng).
17. M. Helal, E. Hingant, L. Pujo-Menjouet, and G. F. Webb, Alzheimer's disease: analysis of a
mathematical model incorporating the role of prions, J. Math. Biol. 69 (2013), no. 5, 1–29.
18. M. D. Hurd, P. Martorell, A. Delavande, K. J. Mullen, and K. M. Langa, Monetary costs of
dementia in the United States, New Engl. J. Med. 368 (2013), no. 14, 1326–1334.
19. J. J. Iliff, M. Wang, Y. Liao, B. A. Plogg, W. Peng, G. A. Gundersen, H. Benveniste, G. E.
Vates, R. Deane, S. A. Goldman, E. A. Nagelhus, and M. Nedergaard, A paravascular pathway
facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes,
including amyloid β, Sci. Transl. Med. 4 (2012), no. 147, 147ra111.
20. C. R. Jack Jr., D. S. Knopman, W. J. Jagust, L. M. Shaw, P. S. Aisen, M. W. Weiner,
R. C. Petersen, and J. Q. Trojanowski, Hypothetical model of dynamic biomarkers of the
Alzheimer's pathological cascade, Lancet Neurol. 9 (2010), no. 1, 119–128.
21.
, Tracking pathophysiological processes in Alzheimer's disease: an updated hypothetical
model of dynamic biomarkers, Lancet Neurol. 12 (2013), no. 2, 207–216.
18
M. BERTSCH, B. FRANCHI, N. MARCELLO, M. C. TESI, AND A. TOSIN
22. E. Karran, M. Mercken, and B. De Strooper, The amyloid cascade hypothesis for Alzheimer's
disease: an appraisal for the development of therapeutics, Nat Rev. Drug Discov. 10 (2011),
no. 9, 698–712.
23. P. Lauren¸cot and S. Mischler, Global existence for the discrete diffusive coagulation-
fragmentation equations in l1, Rev. Mat. Iberoamericana 18 (2002), no. 3, 731–745.
24. M. P. Mattson, Pathways towards and away from Alzheimer's disease, Nature 430 (2004),
631–639.
25. M Meyer-Luehmann, TL Spires-Jones, C Prada, M Garcia-Alloza, A De Calignon,
A. Rozkalne, J. Koenigsknecht-Talboo, D. M. Holtzman, B. J. Bacskai, and B. T. Hyman,
Rapid appearance and local toxicity of amyloid-β plaques in a mouse model of Alzheimer's
disease, Nature 451 (2008), no. 7179, 720–724.
26. J. C. Miller, Neuroimaging for dementia and Alzheimer's disease, Radiology Rounds 4 (2006),
no. 4, 1–4.
27. L. Mosconi, V. Berti, L. Glodzik, A. Pupi, S. De Santi, and MJ. de Leon, Pre-clinical detection
of Alzheimer's disease using FDG-PET, with or without amyloid imaging, J. Alzheimer's Dis.
20(3) (2010), 843–854.
28. R. M. Murphy and M. M. Pallitto, Probing the kinetics of β-amyloid self-association, Bio-
physical Journal 130 (2000), no. 2-3, 109–122.
29. Charles Nicholson and Eva Sykov´a, Extracellular space structure revealed by diffusion analysis,
Trends in Neurosciences 21 (1998), no. 5, 207 – 215.
30. A. Nordberg, Amyloid imaging in Alzheimer's disease, Neuropsychologia 46 (2008), no. 6,
1636–1641.
31. Kenjiro Ono, Margaret M. Condron, and David B. Teplow, Structure-neurotoxicity relation-
ships of amyloid β-protein oligomers, P. Natl. Acad. Sci. USA 106 (2009), no. 35, 14745–
14750.
32. M. M. Pallitto and R. M. Murphy, A mathematical model of the kinetics of β-amyloid fibril
growth from the denatured state, J. Struct. Biol. 81 (2001), no. 3, 1805–1822.
33. R. C. Petersen, R. O. Roberts, D. S. Knopman, B. F. Boeve, Y. E. Geda, R. J. Ivnik, G. E.
Smith, and C. R. Jack Jr., Mild cognitive impairment: Ten years later, Arch. Neurol. 66
(2009), no. 12, 1447–1455.
34. B. Piccoli and A. Tosin, Time-evolving measures and macroscopic modeling of pedestrian
flow, Arch. Ration. Mech. Anal. 199 (2011), no. 3, 707–738.
35. R. A. Quinlan and B. Straughan, Decay bounds in a model for aggregation of microglia:
application to Alzheimer's disease senile plaques, Proc. R. Soc. Lond. Ser. A Math. Phys.
Eng. Sci. 461 (2005), no. 2061, 2887–2897.
36. A. Raj, A. Kuceyeski, and M. Weiner, A network diffusion model of disease progression in
dementia, Neuron 73 (2012), no. 6, 1204–1215.
37. E. M. Reiman, Y. T. Quiroz, A. S. Fleisher, K. Chen, C. Velez-Pardo, M. Jimenez-Del-Rio,
A. M. Fagan, A. R. Shah, S. Alvarez, A. Arbelaez, M. Giraldo, N. Acosta-Baena, R. A. Sper-
ling, B. Dickerson, C. E. Stern, V. Tirado, C. Munoz, R. A. Reiman, M. J. Huentelman, G. E.
Alexander, J. B. S. Langbaum, K. S. Kosik, P. N. Tariot, and F. Lopera, Brain imaging and
fluid biomarker analysis in young adults at genetic risk for autosomal dominant Alzheimer's
disease in the presenilin 1 E280A kindred: a case-control study, Lancet Neurol. 11 (2012),
no. 12, 1048–1056.
38. C. Reitz, C. Brayne, and R. Mayeux, Epidemiology of Alzheimer disease, Nat. Rev. Neurol.
7 (2011), 137–152.
39. J.-M. Serot, J. Zmudka, and P. Jouanny, A possible role for CSF turnover and choroid plexus
in the pathogenesis of late onset Alzheimer's disease, J. Alzheimer's Dis. 30 (2012), no. 1,
17–26.
40. S. J. Tomsky and R. M. Murphy, Kinetics of aggregation of synthetic β-amyloid peptide, Arch.
Biochem. Biophys. 294 (1992), no. 2, 630–638.
41. A. Tosin and P. Frasca, Existence and approximation of probability measure solutions to
models of collective behaviors, Netw. Heterog. Media 6 (2011), no. 3, 561–596.
42. B. Urbanc, L. Cruz, S. V. Buldyrev, S. Havlin, M. C. Irizarry, H. E. Stanley, and B. T.
Hyman, Dynamics of plaque formation in Alzheimer's disease, Biophys. J. 76 (1999), no. 3,
1330–1334.
43. D. M. Walsh and D. J. Selkoe, Aβ oligomers: a decade of discovery, J. Neurochem. 101
(2007), no. 5, 1172–1184.
44. D. Wrzosek, Existence of solutions for the discrete coagulation-fragmentation model with
diffusion, Topol. Methods Nonlinear Anal. 9 (1997), no. 2, 279–296.
45. Wai-Ying Wendy Yau, Dana L Tudorascu, Eric M McDade, Snezana Ikonomovic, Jeffrey A
James, Davneet Minhas, Wenzhu Mowrey, Lei K Sheu, Beth E Snitz, Lisa Weissfeld, Peter J
ALZHEIMER'S DISEASE: A MATHEMATICAL MODEL FOR ONSET AND PROGRESSION19
Gianaros, Howard J Aizenstein, Julie C Price, Chester A Mathis, Oscar L Lopez, and Wil-
liam E Klunk, Longitudinal assessment of neuroimaging and clinical markers in autosomal
dominant alzheimer's disease: a prospective cohort study, The Lancet Neurology 14 (2015),
no. 8, 804 – 813.
Dipartimento di Matematica, Universit`a di Roma "Tor Vergata", Via della Ricerca
Scientifica 1, 00133 Roma, Italy
Istituto per le Applicazoni del Calcolo "M. Picone", Consiglio Nazionale delle
Ricerche, Via dei Taurini 19, 00185 Roma, Italy
E-mail address: [email protected]
University of Bologna, Department of Mathematics, Piazza di Porta S. Donato 5,
40126 Bologna, Italy
E-mail address: [email protected]
S.C. Neurologia, Azienda Ospedaliera Santa Maria Nuova-IRCCS, Viale Risorgi-
mento 80, 42123 Reggio Emilia, Italy
E-mail address: [email protected]
University of Bologna, Department of Mathematics, Piazza di Porta S. Donato 5,
40126 Bologna, Italy
E-mail address: [email protected]
Istituto per le Applicazioni del Calcolo "M. Picone", Consiglio Nazionale delle
Ricerche, Via dei Taurini 19, 00185 Roma, Italy
E-mail address: [email protected]
|
1601.03655 | 2 | 1601 | 2017-02-22T14:10:22 | Shape regulation generates elastic interaction between living cells | [
"physics.bio-ph",
"cond-mat.soft"
] | The organization of live cells to tissues is associated with the mechanical interaction between cells, which is mediated through their elastic environment. We model cells as spherical active force dipoles surrounded by an infinite elastic matrix, and analytically evaluate the interaction energy for different scenarios of their regulatory behavior. We obtain attraction for homeostatic (set point) forces and repulsion for homeostatic displacements. When the translational motion of the cells is regulated, the interaction energy decays with distance as $1/d^4$, while when it is not regulated the energy decays as $1/d^6$. This arises from the same reasons as the van der Waals interaction between induced electric dipoles. | physics.bio-ph | physics |
Shape regulation generates elastic interaction between living cells
Roman Golkov and Yair Shokef∗
School of Mechanical Engineering and The Sackler Center for Computational
Molecular and Materials Science, Tel Aviv University, Tel Aviv 69978, Israel.
The organization of live cells to tissues is associated with the mechanical interaction between
cells, which is mediated through their elastic environment. We model cells as spherical active force
dipoles surrounded by an infinite elastic matrix, and analytically evaluate the interaction energy for
different scenarios of their regulatory behavior. We obtain attraction for homeostatic (set point)
forces and repulsion for homeostatic displacements. When the translational motion of the cells is
regulated, the interaction energy decays with distance as 1/d4, while when it is not regulated the
energy decays as 1/d6. This arises from the same reasons as the van der Waals interaction between
induced electric dipoles.
I.
INTRODUCTION
A. Mechanobiological background
Live cells exert contractile forces on their environment [1]. The elastic behavior of the extracellular matrix (ECM)
and its rigidity affect the forces that cells apply and consequently the resultant displacements [2 -- 5], the cell migratory
behavior [6 -- 10], and cell division [11]. Alterations of the shape and size [12] and also of the rigidity [13] of cells in
response to changes of the ECM rigidity were also observed experimentally. For stem cells the rigidity of the ECM
may even affect their biological phenotype, from neuronal on soft substrates, muscle on substrates with intermediate
rigidity to bone on rigid substrates [14]. It was also shown experimentally that the interaction behavior between cells
is related to the elastic moduli and the non-linear elastic behavior of the substrate [6], and that presence of rigid
boundaries affects the behavior of cells [15].
One of the immediate conclusions from these experimental results is that cells sense changes in their mechanical
environment and respond to those changes in a variety of ways: by modifying the applied forces and displacements,
their size and shape and even their rigidity. This, in turn suggests that there are mechanical interactions between cells
∗ [email protected]
2
through the environment [16]. Indeed, the dependence of the interaction distance between live cells on the rigidity of
the substrate was observed [6]. It was also shown that on nonlinear substrates, cells respond to the presence of each
other over relatively long distances (up to 10 cell diameters) [2] in contrast to linear substrates, on which the response
distance is limited to 1-2 cell diameters. This fact is attributed to the strain-stiffening behavior of the ECM [2, 17] or
to fiber buckling [9]. Improved models of biopolymer gels predict an increase in the range of transmission of internal
cellular forces [18, 19]. An alternative hypothesis is that the fibrous nature of the ECM makes the transmission of
mechanical signals to such long distances possible [20 -- 24].
In this paper we model cells as spherical force dipoles, which create contractile forces on their surface. We seek
the interaction energy between two such cells surrounded by an infinite ECM. Realistically, this ECM has nonlinear
material properties, thus theoretically analyzing such interactions is a very ambitious goal [25, 26]. Here we lay the
conceptual foundations for reaching this goal by focusing on active force dipoles surrounded by a linearly-elastic, or
Hookean material. We suggest that the concepts we introduce and the physical mechanisms that we identify would
be relevant also to nonlinear media. We define the interaction energy as the additional amount of work performed
by the force dipoles as a result of the presence of neighboring dipoles. From studies of the micromechanics of elastic
inclusions we know that in the case of two bodies of any shape that apply a hydrostatic pressure on a surrounding
linearly elastic material, their interaction energy vanishes [27, 28]. We define this type of force dipoles as "dead".
As an example of such behavior one may think of heating a system with inclusions in a medium with a different
thermal expansion coefficient. The self-displacements applied by each such inclusion do not depend on the distance to
neighboring inclusions. In contrast to this, in this paper we introduce "live" behavior as self-regulation of the applied
forces or displacements in order to preserve their shape in the presence of the interaction with other cells, or force
dipoles. We calculate the interaction energy between two "live" force dipoles as a function of the separation distance
between them.
The mechanical environment of the cell may change due to the presence of other cells, due to external forcing, or
due to changes in the medium's rigidity, and it is not clear how cells respond to such changes. The working hypothesis
of many studies, and which we will adopt here, is that there is some sort of mechanical homeostasis in the cell.
Namely, the cell tends to maintain certain quantities [29, 30]. For instance, cells may regulate the forces that they
apply, and then the displacement that they generate will vary with environmental changes, or alternatively, cells may
regulate their deformation, and then the forces required to generate those displacements will vary. We show that
shape regulation of the force dipoles leads to attractive interaction for force homeostasis and to repulsive interaction
3
for displacement homeostasis. The interactions we find are analogous to the van der Waals interaction between two
induced electric dipoles, thus corroborating the mechanobiological elasticity-electrostatics analogy [31 -- 33].
It is important to emphasize that real cells probably do not keep their exact shape, however we suggest that gener-
ically self-regulation related to the interplay between active forces and the cell's shape could generate an interaction
which is qualitatively similar to what we find. Our work deals with an abstract model of the cell and its mechanical
behavior, which in real life are clearly more complicated. Yet we hope that insights obtained from our analytical solu-
tion of this ideal picture will shed light on the understanding of interactions in more realistic scenarios. In particular
it would be interesting to relate our work to recent work on the relation between cell morphology and the polarization
of the active forces that the cell applies in response to its mechanical environment [5, 34, 35].
B. Cells as spherical active force dipoles
In analogy to Eshelby's elastic inclusions as force-generating centers [36, 37], the contractile activity of cells can be
modeled as contractile force dipoles [31], see Fig. 1. We adopt this approach, and for simplicity introduce spherical
force dipoles as spherical bodies, which apply isotropic contractile forces on their environment [25]. The contractile
forces at the surface of each spherical force dipole represent the cellular forces that are generated by the contraction of
the actin network by myosin motors and are transferred to the ECM trough focal adhesions. By analogy with electric
dipoles in electrostatics, a force dipole is defined in mechanics as two equal and opposite point forces applied at some
distance from each other, see Fig. 1(a). Our spherical force dipole consists of an infinite number of such linear force
dipoles with the same center point, distributed isotropically on the surface of the sphere, see Fig. 1(b).
(a)
(b)
FIG. 1: (a) Linear and (b) spherical force dipoles.
C. No interaction between "dead" spherical force dipoles
4
The interaction energy vanishes for two spherical inclusions in an infinite, linearly-elastic medium, each of which
induces symmetric displacements or stresses in the principal directions; namely an initial volume change only with no
distortion of their shapes [27, 28], see Fig. 2. This result may be explained in the following way: the dilation of each
sphere creates a pure shear field around it and a solely compressive field inside. The energy of interaction between
the two spheres may be expressed in terms of the stress from one sphere times the strain from the other sphere,
integrated over the interior of the spheres [28, 38]. Since one field is a pure shear and the other is pure compression,
their coupling does not generate an additional energy. Explicit evaluation of the vanishing interaction energy in this
case [39] is given in Appendix A. We note that this result holds if the elastic moduli inside the inclusions are identical
to those of their environment, while if the elastic moduli differ, the interaction energy will not vanish [40].
r1
d
θ1
R0
u0
1
r2
θ2
θ′
2
u0
2
FIG. 2: Two spherical force dipoles each with radius R0, both applying a radial isotropic displacement u0 on their surfaces.
The coordinate system of sphere 1 is right handed (green) and the coordinate system of sphere 2 is left handed (blue) and may
be written as θ2 = π − θ′
2 where θ′
2 is the commonly-used right-handed azimuthal coordinate for sphere 2.
For calculating the interaction energy in an infinite periodic array of spherical active force dipoles, we first note
that this situation is equivalent to a single force dipole in its Wigner-Seitz unit cell, see Fig. 3. The symmetry of
this periodic array dictates that the normal displacement on the midplane between each two neighboring force dipoles
must vanish, and thus we may equivalently analyze a spherical force dipole surrounded by elastic ECM limited by
a rigid polyhedral Wigner-Seitz unit cell. Since we describe the material as linearly elastic, if we do not take into
account regulation in the activity of the force dipoles due to the deformations generated by their neighbors, we can
employ the aforementioned result regarding two "dead" force dipoles and conclude that also for a system of many
such force dipoles the interaction energy vanishes.
5
FIG. 3: Triangular lattice of spherical force dipoles, its Wigner-Seitz unit cell (blue dashed line), and approximated spherical
unit cell (red dashed line).
D. Spherical Wigner-Seitz unit cell approximation
Instead of solving the displacement field with the boundary condition of vanishing displacement on the actual
boundaries of the polyhedral Wigner-Seitz unit cell, we previously approximated the unit cell as having a spherical
shape [30]. Within this approximation we could exactly calculate the interaction energy by solving the spherically-
symmetric equations of mechanical equilibrium. We define the interaction energy as the difference between the work
performed by the spherical force dipole to displace the medium by u0 when it is surrounded by a rigid spherical
Wigner-Seitz unit cell and the work when it is surrounded by an infinite ECM. For a force dipole of radius R0
surrounded by ECM with bulk modulus K and shear modulus G. The self-energy of a single force dipole in an infinite
medium is given by (see Section II E below): E0 = 8πGR0u2
0. This is the basic energy scale in our problem, and all
interaction energies we obtain scale with E0. For a single force dipole enclosed by a spherical Wigner-Seitz unit cell
of radius Rc, we relate Rc to the distance between neighboring force dipoles, and define the dimensionless distance
as Rc = Rc/R0. The interaction energy in the displacement-homeostasis scenario is given by [30]: Eint = 3−ν
4(1−ν)
E0
eR3
c
,
where ν = 3K−2G
2(3K+G) is the Poisson ratio of the ECM. This interaction energy represents a repulsive force which decays
algebraically with dipole-dipole separation. If instead we consider stress homeostasis and fix the stress σ0 that the
force dipole applies on the medium, we may write E0 = πσ2
0 R3
0
2G and now Eint = − 3−ν
4(1−ν)
E0
eR3
c
, representing an attractive
interaction.
E. Motivation
6
At first glance it seems that these results [30] contradict the previously-mentioned general result [27, 28] that bodies
of any shape included in a continuous linear elastic material and applying on it isotropic dilational displacements on
their surface do not interact. In order to resolve this seeming contradiction, in this paper we compare the case of
a spherical force dipole inside a concentric spherical Wigner-Seitz unit cell (see Fig. 3) to the realistic configuration
in which two spherical force dipoles are embedded at some distance d between them in an infinite ECM with linear
properties (see Fig. 2). This comparison enables us to demonstrate that the reason for the contradiction is the
difference between the definitions of the geometries in the two cases; In the two-sphere case the force dipoles obtain
a drop-like shape due to the interaction [see Fig. 4(a)], while a spherical force dipole inside a spherical Wigner-Seitz
unit cell preserves its spherical shape even when it interacts with other force dipoles. Since the displacements created
by each force dipole distort its neighbor, shape preservation may be achieved in the two-force-dipole setup only by
application of appropriate anisotropic displacements by each dipole on its surface [see Fig. 4(b),4(c)], and these give
rise to the interaction energy that we study in this paper.
II. METHODS
A. Theoretical framework
We consider two identical spheres of radius R0 surrounded by an infinite material with linear elastic, or Hookean
behavior defined by bulk modulus K and shear modulus G and denote the distance between their centers by d. In
addition to the isotropic displacement u0, each sphere applies an anisotropic displacement, which is intended to cancel
the anisotropic displacements on its surface that are caused by the other sphere. In order to simplify the calculations
we choose a left-handed coordinate system for sphere 2; namely the newly defined angle θ2 equals π − θ′
2, see Fig. 2.
B. General displacements generated by spherical force dipoles
The displacement field around each force dipole must satisfy mechanical equilibrium, which we write in terms of
the displacements field ~u as [41]:
1
1 − 2ν ∇∇ · −→u + ∇2−→u = 0.
(1)
7
(a)"Dead"
(b)"Live", fixed-size, fixed-position
(c)"Live", variable-size, variable-position
FIG. 4: Two contractile spherical force dipoles in an infinite elastic medium: (a) "Dead" force dipoles applying an isotropic
elastic force, (b-c) "Live" force dipoles regulating the force they apply in order to remain spherical even in the presence of the
other force dipole. Red arrows are forces applied by each force dipole, black dashed lines are the corresponding self displacements,
blue dotted lines are the displacements caused by the other force dipole, and green solid line is the total displacement. For
illustration purposes, the initial distance between the spheres was set to d = 3R0, and the self-displacement to u0 = 0.4R0.
The Poisson ratio is ν=0.05. We show two of the "live" cases that we describe below, which differ in whether the volumes and
positions of the spheres are preserved during the interaction (b) or not (c).
Due to the symmetry with respect to rotation about the axis connecting the centers of the two force dipoles, there is
no dependence on the azimuthal angle φ, thus we write Eq. (1) in spherical coordinates as:
8
1
1 − 2ν
1
1
r
1 − 2ν
r2
∂
∂r(cid:20) 1
∂θ(cid:20) 1
∂
r2
∂
∂r
∂
∂r
(r2ur) +
(r2ur) +
1
r sin θ
1
r sin θ
∂
∂θ
∂
∂θ
(uθ sin θ)(cid:21) + ∇2ur −
(uθ sin θ)(cid:21) + ∇2uθ +
2
r2
2
r2 ur −
∂ur
2
∂θ −
r2
∂uθ
∂θ −
uθ
= 0,
2uθ cot θ
r2
= 0.
r2 sin2 θ
where the Laplacian in spherical coordinates excluding terms depending on φ is given by:
∇2 =
1
r2 sin θ(cid:20) ∂
∂r(cid:18)r2 sin θ
∂
∂r(cid:19) +
∂
∂θ(cid:18)sin θ
∂
∂θ(cid:19)(cid:21) .
(2)
(3)
(4)
Based on the general solution for the displacement field of a sphere with given cylindrically symmetric displacements
on its surface [41], we write the anisotropic displacements field satisfying (2-3) outside the spherical force dipole
(r > R0) as a multipole expansion in terms of spherical harmonics Yn(θ) =q 2n+1
4π Pn(cos θ):
uri =
+ u0
u0R2
0
r2
i
∞Xn=0(cid:20)n(n + 3 − 4ν)
uθi = u0
k=0 xk
n
k
∞Xn=0(cid:20)(−n + 4 − 4ν)
n+k−1
n
2
DnRn+2
0
rn+2
i
CnRn
0
i − (n + 1)
rn
(cid:21) dYn(θi)
DnRn+2
0
rn+2
i
dθi
,
CnRn
0
rn
i
+
(cid:21) Yn(θi),
(5)
(6)
the Legendre polynomial of order n [42]. Here uri and uθi are the radial
and angular components of the displacement field caused by sphere i, and the infinite sums represent the anisotropic
corrections that each force dipole produces in order to cancel the shape distortion caused by its neighbor. u0 and R0
have been inserted so that the coefficients Cn and Dn are dimensionless.
Defining the dimensionless displacementseuri = uri
dimensionless form as:
and positioner = r
R0
, we rewrite Eqs. (5) and (6) in
with Pn(x) = 2n·Pn
1
u0
u0
+
, euθi = uθi
∞Xn=0(cid:20)n(n + 3 − 4ν)
ern
er2
∞Xn=0(cid:20)(−n + 4 − 4ν)
Cn
i
i
+
euri(ri, θi) =
euθi(ri, θi) =
Dn
Cn
i − (n + 1)
ern+2
(cid:21) dYn(θi)
Dn
dθ
.
i
ern
ern+2
i
(cid:21) Yn(θi),
(7)
(8)
Note that (7-8) solve Eq. (1) only when each force dipole is surrounded by an infinite homogeneous linearly-elastic
medium, including in the interior of the neighboring force dipole. Biological cells clearly have a rigidity which differs
from that of the ECM surrounding them, and thus this assumption seems to be problematic. We overcome this by
realizing that we may first solve the mechanical problem in which the cells are assumed to have the same linear elastic
properties as the ECM. The resultant solution includes a certain stress and displacement on the surface of each cell,
and the solution outside the cells is independent on how the cell generates this stress on its surface. In particular, the
stress that actual cells apply on their surrounding includes a passive stress coming from the rigidity of the cell plus
an active stress coming from the external forces generated by molecular motors inside the cell. In our analysis we
consider only the total stress and the work performed by it, which determines the interaction energy, and our results
are valid irrespective of the mechanical rigidity of the cells themselves.
9
C. Cancellation condition
For our "live" force dipoles the sum of the anisotropic displacements caused by the neighbor force dipole and of
all the corrections applied by the discussed force dipole must vanish on its surface. From this we derive conditions
for the coefficients Cn and Dn so that each force dipole will preserve its spherical shape even in the presence of the
interaction with its neighboring force dipole. In order to apply the cancellation condition and to derive from it the
expressions for Cn and Dn, we transform the expressions for the displacement field of force dipole 2 to the coordinate
system of force dipole 1 by substitution of the expressions for r2 and θ2 in terms of r1 and θ1 and then multiplying
the displacement vector −→u2 = (ur2, uθ2) by the rotation matrix:
B =
− cos (θ1 + θ2) sin (θ1 + θ2)
cos (θ1 + θ2)
sin (θ1 + θ2)
.
(9)
(10)
(11)
We then write the resultant expressions for the radial and angular displacements caused by force dipole 2 on the
surface of force dipole 1 in terms of the spherical harmonics of sphere 1 by writing:
(ur)n =
(uθ)n =
2
2n + 1
Z π
2n(n + 1)Z π
2n + 1
0
0
ur(θ)Yn(θ)sinθdθ,
uθ(θ)
Yn(θ)
dθ
sinθdθ.
As may be seen from (10,11), every spherical-harmonic mode of force dipole 2 contributes to all the modes on the
surface of force dipole 1.
As explained above, the sum of the anisotropic displacements caused by force dipole 2 must be canceled on the
surface of force dipole 1 by the corrections that it applies. We write the dimensionless displacement eu11 created by
force dipole 1 on its surface (namely ater1 = 1):
∞Xn=0hn(n + 3 − 4ν)Cn − (n + 1)(Dn −
∞Xn=0
√4πδn,0)i Yn(θ1),
eur11(θ1) ≡eur1(1, θ1) =
euθ11(θ1) ≡euθ1(1, θ1) =
[(−n + 4 − 4ν)Cn + Dn]
dYn(θi)
(12)
(13)
.
dθ
The term δn,0 in (12) is a Kronecker delta, which represents the isotropic radial displacement created by sphere 1 on
its surface without the anisotropic cancellation corrections. This constant term does not depend on changes in the
environment of the force dipole. The remaining terms are different modes of additional displacement created by this
"live" force dipole in response to the displacement field induced on its surface by the neighboring force dipole. The
10
dimensionless displacementeu21 created by force dipole 2 on the surface of force dipole 1 is:
√4πδm,0)i Yn(θ1),
√4πδm,0)i dYn(θ1)
∞Xm=0hf Cr
∞Xm=0hf Cθ
nm(ed)Cm + f Dr
nm(ed)Cm + f Dθ
nm(ed)(Dm −
nm(ed)(Dm −
eur21(θ1) =
euθ21(θ1) =
∞Xn=0
∞Xn=0
dθ1
(14)
(15)
,
where the sum over m originates from the fact that the displacement eu2 created by force dipole 2 is given by a
multipole expansion (7-8) with the corrective magnitudes Cm and Dm. The sum over n originates from the fact
that after the coordinate transformation, when these modes are expressed in terms of the spherical harmonics in the
coordinate system of force dipole 1, each mode from force dipole 2 contributes to all the modes of force dipole 1.
R0
The functions f Cr
between the spherical force dipoles.
nm(ed), f Cθ
nm(ed), f Dr
nm(ed) and f Dθ
nm(ed), given in Appendix B depend only on the dimensionless distance
ed = d
We now require that for "live" force dipoles the total displacementeu11(θ1) +eu21(θ1) on the surface of force dipole
1 is isotropic. We begin by considering the simplest (but strictest) regulation scenario, for which not only is this total
displacement isotropic, but its magnitude remains equal to the displacement u0 in the absence of interactions between
the force dipoles. Moreover, we require that the center of symmetry of each force dipole does not move. We will later
consider three additional scenarios in which the interaction causes the force dipoles to change their volume and/or to
move, yet they remain spherically symmetric. Thus at this point we require that:
Substituting (12,13,14,15) in (16,17) yields:
eur11(θ1) +eur21(θ1) ≡ 1,
euθ11(θ1) +euθ21(θ1) ≡ 0.
(16)
(17)
∞Xn=0(hn(n + 3 − 4ν)Cn − (n + 1)(Dn −
∞Xn=0([(−n + 4 − 4ν)Cn + Dn] +
∞Xm=0hf Cθ
√4πδn,0)i +
nm(ed)Cm + f Dθ
∞Xm=0hf Cr
nm(ed)(Dm −
nm(ed)Cm + f Dr
nm(ed)(Dm −
√4πδm,0)i) dYn(θ1)
dθ1
√4πδm,0)i) Yn(θ1) = 1, (18)
= 0.
(19)
Due to the orthogonality of the Legendre polynomials, for these infinite sums to satisfy these cancellation conditions,
each term in the sums must cancel independently. Thus for all n ≥ 1 we require:
n(n + 3 − 4ν)Cn − (n + 1)Dn +
nmCm + f Dr
nm(Dm −
(−n + 4 − 4ν)Cn + Dn +
nmCm + f Dθ
nm(Dm −
∞Xm=0hf Cr
∞Xm=0hf Cθ
√4πδm,0)i = 0,
√4πδm,0)i = 0.
11
(20)
(21)
Note that for n = 0, from (7-8) C0 is irrelevant, thus we set it to zero. Moreover, since Y0(θ1) = 1, dY0(θ1)
dθ1
= 0 and
(19) holds trivially, thus for n = 0 we obtain only one equation, from Eq. (18):
−D0 +
∞Xm=0hf Cr
0mCm + f Dr
0m (Dm −
√4πδm,0)i = 1.
(22)
We obtain closure of the infinite coupled linear Eqs. (20, 21, 22) by assuming that Cn = 0 and Dn = 0 for n > nmax.
This is justified since we will be interested in large separations between the force dipoles, and due to the fact that the
solutions decay as 1/rn, thus at large r, large n terms become negligible. We verify this numerically by increasing
nmax until convergence, see Section III below.
D. The n=0,1 modes
Each of the spherical harmonics Yn(θ) represents a different mode of deformation of the force dipole. For scalar fields
the first two modes represent volume change (n = 0) and translation of the center of symmetry (n = 1). However, as
can be seen from Eqs. (7,8), since the displacement field is vectorial, here the n = 1 mode may represent a combination
of translation and deformation. Since our definition of "live" force dipoles focuses only on shape regulation, inclusion
in the solution of the terms n = 0 and the part of the n = 1 term that controls translation is not mandatory. Inclusion
of one or both of these terms will result in complementary condition(s) in addition to the shape preservation condition
and will describe a different type of self-regulation of the force dipoles. Overall there are four combinations for the
inclusion or exclusion of the n = 0 and n = 1 terms and thus we analyze four possible scenarios of homeostasis or
self-regulation, see Fig. 5 and Fig. 4(b)-4(c) above.
As a result of exclusion of one or more of the terms of the sum, Eqs. (20-22) will not vanish anymore at the
corresponding modes, i.e. non-zero resultant displacement will exist. At the same time the force applied by the dipole
and the resulting elastic stress will vanish at these modes. Mode n = 0 of the displacement never vanishes since it
includes the constant term u0 which represents the fixed symmetric displacement created by the force dipole, which
does not depend on the interaction. For the size (n = 0) we have two options - fixed size (FS) namely displacement
homeostasis, in which we regulate this mode, and variable size (VS) namely stress homeostasis, in which we allow it
Fixed Position:
Variable Position:
12
Fixed Size:
Variable Size:
FIG. 5: The four possible scenarios of shape regulation, depending on whether the size (n = 0) and the position (n = 1) are
regulated or not.
to have an additional contribution due to the interaction. Similarly we consider either fixed position (FP) or variable
position (VP) of the force dipole (n = 1). In the VP case instead of the cancellation condition (20-21), the coefficients
C1, D1 obey the following:
(4 − 4ν)C1 − 2D1 +
(f Cr
1mCm + f Dr
1m Dm) =
(3 − 4ν)C1 + D1 +
(f Cθ
1mCm + f Dθ
1m Dm) =
∞Xm=0
∞Xm=0
2
∞Xm=0(cid:18) f Cr
∞Xm=0(cid:18) f Cr
2
1m + f Cθ
1m
f Dr
1m + f Dθ
1m
Cm +
2
1m + f Cθ
1m
Cm +
f Dr
1m + f Dθ
1m
2
Dm(cid:19) ,
Dm(cid:19) .
(23)
(24)
Equations (23-24) allow us to find C1 and D1 that cancel the deformation at mode 1 without canceling the transla-
tional motion of the force dipole. They follow from Eqs. (20-21) and from the fact that for translation the coefficients
of P1(cosθ) = cos(θ) in the radial direction and of dP1(cosθ)
dθ
= sin(θ) in the tangential direction must be equal.
E.
Interaction energy
After obtaining the coefficients Cn and Dn and substituting them in Eqs. (12-15) we compute the interaction energy
from the work performed by the force dipoles to generate this deformation:
Here the integration is over the surfaces of both spheres, −→F is the force per unit area applied by each force dipole on
∆E = E − 2E0 =
1
2ZS(cid:16)−→u · −→F − −→u0 · −→F0(cid:17) ds.
(25)
its environment. It is important to emphasize the difference in calculation of the interaction energies in the cases of
dead versus live force dipoles. In the general case of two dead force dipoles the displacements and forces applied by
each one of them are not affected by changes in its environment and thus the amount of additional work done by it
equals [43] [see Eq. (25)]:
∆Ei =
1
2ZSih(−→ui + −→uj) · −→Fi − −→ui · −→Fii dSi =
1
2ZSi
−→uj · −→FidSi.
(26)
In the particular case that the force dipoles apply isotropic forces or displacements causing only volume change without
distortion (as for two "dead" spherical force dipoles) the interaction energy ∆E vanishes [27]. In contrast, for two
"live" force dipoles the forces and displacements created by each one of them are modified due to the interaction
with the neighbor. The resultant forces and displacements applied by "live" force dipoles are not isotropic and the
interaction energy does not vanish anymore. Here the additional work that is performed by each force dipole is:
13
∆Ei =
=
1
2ZSih(−→ui + δ−→ui + −→uj + δ−→uj)(−→Fi + δ−→Fi) − −→ui · −→Fii dSi
2ZSih−→uj · −→Fi + δ−→ui · −→Fi + δ−→uj · −→Fi + −→ui · δ−→Fi + −→uj · δ−→Fi + δ−→ui · δ−→Fi + δ−→uj · δ−→Fii dSi.
1
(27)
Here δ−→u and δ−→F are the "live" parts of the displacement and the force created by the force dipoles due to their
interaction. For two "live" force dipoles, we evaluate the interaction energy using the stress tensor τ that arises in
the elastic environment of each force dipole in response to the displacement −→u on its surface. The energy evaluated
in Eq. (27) equals the work done by spherical force dipole i on its surface and thus we add the index i and the total
interaction energy is E1 + E2. The forces applied by each force dipole are equal and opposite to the forces applied on
it by the environment: −→F = −τ · r, where r is the outward pointing unit vector normal to the surface of the sphere
after its deformation and movement. Since we allow only volume change and translational motion of the entire force
dipoles, the direction of the normal does not change due to the interaction. The terms of the stress tensor τ based on
the general solution of the displacements field Eqs. (5-6) [41] are given in Appendix C. −→F0 = 4Gu0
unit area on the surface of the sphere in the case of a single force dipole with known isotropic displacement −→u0 = u0r
r is the force per
R0
on its surface and without interactions with other spheres, and thus the self energy is E0 = 8πGu2
0R0.
Evaluation of the integral in Eq. (25) may be simplified by noting that both the displacements −→u and the stresses τ
are expressed in terms of Legendre polynomials and their derivatives [see Eqs. (20-21)] and using their orthogonality
we get (see Appendix D):
E = 2 ·
1
2ZS
−→u · −→F ds = 2E0 1 + D0 −(−D0 +
∞Xm=0hf Cr
0mCm + f Dr
0m (Dm −
√4πδm,0)i)! .
In the FS cases D0 is evaluated from the cancellation condition and Eq. (28) becomes:
EF S = 2E0 (1 + D0) = 2E0 1 −( ∞Xm=0hf Cr
0mCm + f Dr
0m(cid:16)Dm −
√4πδm,0(cid:17)i)! ,
while in the VS cases D0 is zero and Eq. (28) becomes:
EV S = 2E0(1 +
∞Xm=0hf Cr
0mCm + f Dr
0m(cid:16)Dm −
√4πδm,0(cid:17)i) .
(28)
(29)
(30)
The meaning of the sum of the last two terms in Eqs. (29-30) is the change of the volume of force dipole 1 caused by
all the modes of displacement created by force dipole 2 (see Fig. 2). We thus define:
14
∆V21 ≡
∞Xm=0hf Cr
0mCm + f Dr
0m(cid:16)Dm −
Thus the interaction energies in the FS cases become:
and in the VS cases:
∆EF S = −2E0∆V21,
∆EV S = 2E0∆V21.
III. RESULTS
√4πδm,0(cid:17)i .
(31)
(32)
(33)
We evaluated the interaction energy by terminating the infinite sums (20-21) at different values of nmax, up to
nmax = 15 for all four regulation scenarios. Figure 6 shows the normalized interaction energy ∆ E = ∆E/E0
vs the normalized distance between the force dipoles d = d/R0 for all four regulation scenarios. Figure 7 shows the
convergence of the interaction energy with increasing the value of nmax. Even at the smallest dipole-dipole separations
the solution converges with moderate value of nmax = 4 − 5. Moreover, for distances larger than 4 sphere radii in
the FSVP, VSFP and VSVP cases and for distances larger than 6 sphere radii in the FSFP case the energy is well
0.01
4
10
6
10
8
10
E(cid:2)
(cid:1)
a
FSFP
1
2
14
FSVP
2
3
15
0.01
4
10
6
10
8
10
E(cid:3)
(cid:2)
(cid:1)
b
VSFP
1
2
14
VSVP
2
3
15
2
5
10
20
50
100
2
5
10
20
50
100
d(cid:1)
d(cid:1)
FIG. 6: Dimensionless interaction energy ∆ eE = ∆E
E0
vs dimensionless distance between the two force dipoles ed = d
R0
, with
different values of nmax as indicated in the legend. Poisson ratio was set to ν = 0.45. Note that the interaction energy in the
FS cases is positive (repulsive), and in the VS cases negative (attractive). The far-field behavior in the FP cases is given by
1/ d4 and for VP by 1/ d6.
15
FIG. 7: Interaction energy vs inverse of highest term nmax included in the solution. Energy is rescaled by the result with the
minimal possible value of nmax. Results are shown for Poisson ratio ν = 0.45 and dimensionless distance d = 2.5 between the
centers of the two force dipoles.
approximated by taking the minimal nmax possible. Namely for the FP cases nmax = 1 suffices, and for the VP cases
we take nmax = 2, see Table I. Since in Eq. (20) the coefficient of C0 vanishes and Eq. (21) vanishes for n = 0 there
is no need to evaluate the coefficient C0. Thus in the FS cases we get three coupled linear equations and in the VS
cases two coupled linear equations. It is important to reiterate the role of the n = 0 and n = 1 terms in the solution.
Excluding the n = 1 term of the sum in Eqs. (5-6) constitutes the VP cases in which the force dipoles are free to move
but preserve their size and shape, and excluding the n = 0 term constitutes the VS cases in which the force dipoles
preserve their spherical shape but are free to change their volume due to the interaction. We see from Eq. (28) that
the presence of the coefficient D0 determines the sign of the interaction energy, see also Fig. 6 and Table I; Force
dipoles with FS regulation (D0 6= 0), are repelled, while force dipoles with VS regulation (D0 = 0) are attracted.
This is similar to our analysis within the spherical unit-cell approximation of different homeostasis scenarios, where
we found repulsion when the force dipole generates a fixed displacement on its boundary and attraction when it
applies a fixed stress on its environment [30]. Relating this to our present results, in the FS cases the resultant
displacement on the boundary of each force dipole is fixed and does not depend on the changes in its environment,
which corresponds to displacement homeostasis. In the VS cases the size of each force dipole changes due to the
interaction, namely the displacement is not fixed. There is no correction to the n = 0 mode, which means that there
is no additional active force or stress at this mode, and we relate this to stress homeostasis, in which indeed the
interaction is attractive.
Fixed Size
Fixed Size
Variable Size
Variable Size
Fixed Position Variable Position Fixed Position Variable Position
(F SF P )
(F SV P )
(V SF P )
(V SV P )
16
nmax
1
2
(1−2ν)
5−6ν
1
ed4
5(1−2ν)
4−5ν
1
ed6
1
0
2
0
D0
C1
D1
C2
D2
− 3
2(5−6ν)
1
ed2
9(2−3ν)
4(4−5ν)(5−6ν)
1
ed7
− 3
2(5−6ν)
1
ed2
9(2−3ν)
4(4−5ν)(5−6ν)
1
ed7
− 1
2(5−6ν)
1
ed2
9(2−3ν)(7−8ν)
4(4−5ν)(5−6ν)
1
ed7
− 1
2(5−6ν)
1
ed2
9(2−3ν)(7−8ν)
4(4−5ν)(5−6ν)
1
ed7
0
0
− 5
4(4−5ν)
1
ed3
− 3
2(4−5ν)
1
ed3
0
0
− 5
4(4−5ν)
1
ed3
− 3
2(4−5ν)
1
ed3
∆V21 −
2(1−2ν)C1
3 ed2
−
4(1−2ν)C2
ed3
−
2(1−2ν)C1
3 ed2
−
4(1−2ν)C2
ed3
∆ eE∞
2(1−2ν)
(5−6ν)
· 1
ed4
10(1−2ν)
(4−5ν)
· 1
ed6
−
2(1−2ν)
(5−6ν)
· 1
ed4
−
10(1−2ν)
(4−5ν)
· 1
ed6
TABLE I: Coefficients Cn and Dn of the multipole expansion at asymptotically long-distances, ed ≫ 1. Here nmax defines
the highest order term taken into account, and ∆ eE∞ is the asymptotic long-distance behavior of the resultant dimensionless
interaction energy ∆ eE ≡
Eint
E0
. ed = d
R0
is the dimensionless distance between the force dipoles, and ∆V21 is defined in Eq. (33).
The corresponding expressions for arbitrary distance are given in Appendix E.
An intuitive explanation for the sign difference in the interaction may be considered as follows; FS means that
in the proximity to other force dipoles, each force dipole has to exert an additional force in order to generate the
displacement that it was programmed to have. Thus more energy is required and the force dipoles would benefit
energetically from moving away from each other. As noted above, VS is related to stress homeostasis, for which the
tension that neighboring contractile force dipoles generate around themselves add up, and less energy is required in
order to reach the homeostatic stress value, thus it is beneficial for cells to be close to other cells.
The position regulation (n = 1) sets the strength of these attractive or repulsive interactions. For VP, C1 = D1 = 0
and at long distances the interaction energy decays as 1/d6, while for FP, C1 and D1 are nonzero, which leads to a 1/d4
decay, see Fig. 6 and Table I. To quantitatively relate these functional forms to the results of our spherical unit-cell
approximation, we consider an infinite array of active force dipoles. Using our results for the interaction between
two active force dipoles, we now evaluate the interaction energy in an infinite 3D array of such shape-regulating
spherical force dipoles (see Fig. 3). In a linear medium and for small displacements generated by each force dipole,
by superposition the total interaction energy per force dipole is:
Etot =Z ∞
r=0
nEint(r) · 4πr2dr ≈Z ∞
r=D
nEint(r) · 4πr2dr,
17
(34)
where n is the density of force dipoles at distance r from any given force dipole, which for large distances we assume
is uniform, and we introduce a near-field cutoff distance equal to the lattice spacing D.
Substituting the expressions for the two-force-dipole interactions from Table I in Eq. (34) and performing the
integrations, yields: Etot ∝ 1/D (FP), 1/D3 (VP). The VP cases are quantitatively consistent with the results of
our previous work within the spherical unit-cell approximation [30] that the interaction energy in an array of "live"
force dipoles scales as 1/R3
c, since the radius Rc of the periodic unit cell is related to the distance D between force
dipoles in the periodic array. Due to the symmetry of the spherical unit-cell setup, by definition there is no need for
the force dipole to exert any force in order to remain in place and thus there is no regulation of the position, which
corresponds to our VP case.
The expressions for the interaction energy given in Table I (and similarly in Table II in Appendix E for arbitrary
distance) all vanish in the incompressible limit (ν = 0.5). The self-displacement generated by each force dipole
separately includes a pure compressive mode inside that force dipole, and sustaining this in an incompressible medium
with a fixed displacement u0 on the boundary requires an infinite force. Thus the problem of mechanical interactions
between force dipoles in a completely incompressible medium deserves a separate analysis, which we defer to future
work.
IV. DISCUSSION
We model live cells as spherical force dipoles surrounded by a linear, or Hookean elastic environment. Since in
the case of isotropic displacements and forces the interaction energy between force dipoles vanishes, we distinguish
between "dead" behavior in which the force dipoles apply constant forces and self-displacements on their surface, and
"live" behavior in which the forces and self-displacements applied change in response to changes in the environment of
the force dipoles. We solved the interaction energy for four different types of such regulation in which the force dipoles
preserve their spherical shape and in addition volume, position, both or none of them. We found the interaction energy
to be inversely proportional to the distance between the force dipoles to the fourth power in the case of fixed position
and to the sixth power in the case of variable position. These results are similar to and stem from the same reasons
as the van der Waals dipole-induced dipole interaction in electrostatics [44]. We also found that in the case of volume
18
preservation the force dipoles are repelled while without volume preservation they are attracted.
We solved the deformation fields for the case when the rigidity of the force dipoles is identical to that of their
environment. However, biological cells are complex entities and their rigidity differs from point to point and also
differs from the rigidity of the extracellular matrix. In order to relate our results to live cells we describe each of them
as a mechanism that applies forces on its environment on the surface and responds by their variation to application
of external force or displacement. The displacements and forces applied by a cell may be divided into "passive" and
"active" parts. The "passive" part of the forces or displacements would stay the same if the cells were dead and
preserve their elastic properties, while the "active" part depends on the programmed behavior of the cells and is
generated by the contraction of acto-myosin networks inside them. Since the resultant force and displacement are the
sum of those two parts cells may create such "active" response such that the resultant forces and displacements will
coincide with the case considered here, for which their rigidity coincides with the rigidity of their environment.
Stress homeostasis vs displacement homeostasis is believed to occur in different cell types and in different mechanical
environments [45]. We suggest to examine in experiments the correlation that we find here between these distinct
regulatory behaviors and attraction vs repulsion. Specifically, for cells that are known to regulate their stress or
displacement one could test whether they indeed are attracted or repelled, respectively. Alternatively, when the
regulatory behavior is not known, we suggest that our results will enable to infer it from the direction of the interaction.
It would be interesting to consider also more complicated regulatory behaviors on top of the two extreme limits of fixed
stress and fixed displacement. In particular, it would be interesting to test whether our fixed-position vs variable-
position cases could generate more complex scenarios, since we predict that that would relate to the interaction
strength. Clearly the interaction energy that we focus our analysis on may not be directly probed in experiments. We
suggest that our predictions on attraction or repulsion between cells could be tested by studying other experimental
indications of inter-cellular interactions. For instance, cells that are attracted to each other tend to try and move
closer together but in rigid environments may not move and instead send protrusions one toward each other [6].
Moreover, focusing of mechanical stress between interacting cells may be experimentally visualized [2].
Following our work on three-dimensional spherical force dipoles, it would be interesting to solve the case of two-
dimensional disks on an elastic medium. The same four regulatory possibilities can be considered. This could more
directly be related to experiments in which cells are grown on surfaces of different elastic materials [6]. We expect
that in this case the interaction energy will not vanish even for "dead" force dipoles. Clearly the work presented in
this paper is a first step toward theoretical analysis of interactions that stem from mechanobiological regulation, and
in particular shape regulation of live cells. Our focus on linearly-elastic, homogeneous and isotropic media enabled us
to obtain the detailed analytical description presented above. It would be important to continue this line of research
and study the influence of the material nonlinearity, heterogeneity and anisotropy of the ECM on the interaction
between cells.
19
Acknowledgments
We thank Dan Ben-Yaakov, Kinjal Dasbiswas, Haim Diamant, Erez Kaufman, Ayelet Lesman, Sam Safran, Nimrod
Segall, Eial Teomy, Daphne Weihs and Xinpeng Xu for helpful discussions. This work was partially supported by the
Israel Science Foundation grant No. 968/16, by a grant from the United States-Israel Binational Science Foundation
and by a grant from the Ela Kodesz Institute for Medical Physics and Engineering.
Appendix A: Vanishing interaction energy between "dead" cells
Here we calculate the vanishing interaction energy between two "dead" spherical force dipoles each with radius
R0 applying an isotropic radial force F on its environment, see Fig. 2. The environment behaves as a linear elastic
material with shear modulus G. The displacements field around a single such force dipole is given by [46]:
In order to evaluate the resultant displacement of two such force dipoles we superimpose the displacement fields of
both force dipoles. To do so, we translate the displacement field created by force dipole 1 to the coordinate system of
force dipole 2. Using the cosine theorem we rewrite the displacement fields created by the force dipoles in terms of θ2
and d, see Fig. 2:
Here br1 and br2 are unit vectors in the radial directions of force dipoles 1 and 2, respectively. The resultant displacement
in the radial direction of force dipole 2 is:
ur =(cid:20) F R
4G br2 +
4G (R2
F R3
0
0 + d2 + 2R0d cos θ2)br1(cid:21) · br2 =
F R0
4G "1 +
R2
0 (R0 + d cos θ2)
0 + d2 + 2ad cos θ2)3/2# .
(R2
(A4)
−→u (R) =
F R3
0
4Gr2br.
F R3
0
0 + d2 + 2R0d cos θ2)br1,
−→u1 =
−→u2 =
4G (R2
F R0
4G br2.
(A1)
(A2)
(A3)
The symmetry of the system imposes that the total work done by the system reads:
E =Z π
0 (cid:18)2 ·
1
2
F ur · 2πR0 sin θ2(cid:19) R0dθ2 =
After evaluation we obtain:
πF 2R3
0
2G Z π
0 "sin θ2 +
(R0 + d cos θ2) R2
0 + d2 + 2R0d cos θ2)3/2# dθ2.
0 sin θ2
(R2
E =
πF 2R3
0
2G
= 2 · 4πR2
0 ·
1
2
F ·
F R0
4G
= 2E0.
20
(A5)
(A6)
Thus the interaction energy vanishes in this case. A similar analysis may be done for two force dipoles which apply
radial displacements u0 on their surfaces. The energy in this case may be written as:
E = 16R0πu2
0G = 2E0,
(A7)
and also here the interaction energy vanishes.
Appendix B: The functions f Cr
nm, f Dr
nm, f Cθ
nm and f Dθ
nm
The expressions for the functions f Cr
nm, f Dr
nm, f Cθ
nm and f Dθ
nm appearing in Eqs. (18-31) are:
2
2
(2n + 1)
(2n + 1)
Z π
0 (cid:2)g1
Z π
0 (cid:2)g3
2n(n + 1)r 2n + 1
2n(n + 1)r 2n + 1
(2n + 1)
(2n + 1)
4π
f Cr
f Dr
nm(ed) =
nm(ed) =
nm(ed) =
nm(ed) =
f Cθ
f Dθ
4π
Z π
0 (cid:2)g5
Z π
0 (cid:2)g7
=r 2n + 1
4π
We used the identity [47]
mYm (ψ) + g2
mYm (ψ) + g4
mYm+1 (ψ)(cid:3) Yn(θ1)dθ1,
mYm+1 (ψ)(cid:3) Yn(θ1)dθ1,
mYm (ψ) + g6
mYm (ψ) + g8
mYm+1 (ψ)(cid:3) [cos(θ1)Pn(θ1) − Pn+1(θ1)]dθ1,
mYm+1 (ψ)(cid:3) [cos(θ1)Pn(θ1) − Pn+1(θ1)]dθ1.
dYn(θ)
dθ
(n + 1)(cid:20) Pn+1(cos θ)
sin θ
− cot θPn(cos θ)(cid:21)
to rewrite the derivatives dYn(θ)
dθ
in terms of θ, and for the sake of brevity we defined:
g1
m =ned2(cid:2)m2 − m(3 − 4ν) − 4(1 − ν)(cid:3) − 2ed(cid:0)m2 − 2 + 2ν(cid:1) cos(θ1) + m2 + m(3 − 4ν)o sin(θ1)/ζm+1,
g2
m = −hed(m + 1)(m − 4 + 4ν)i sin(θ1)/ζm,
g3
m = − (m + 1) sin(θ1)/ζm+1,
g4
m = ed(m + 1) sin(θ1)/ζm+2,
(B1)
(B2)
(B3)
(B4)
(B5)
(B6)
(B7)
(B8)
(B9)
m = −(cid:16)ed2 + 1(cid:17) (m + 1)(m − 4 + 4ν) cos(θ1) +ed(cid:2)m(m − 6 + 8ν) − 6(1 − ν) +(cid:0)m2 − 2 + 2ν(cid:1) cos(2θ1)(cid:3)
sin(θ1)ζm+2
g5
where
g6
m = −(m + 1)(m − 4 + 4ν)/[sin(θ1)ζm],
g7
m = (m + 1) cot(θ1)/ζm+1,
g8
m = −[ed cos(θ1) − 1](m + 1)/[sin(θ1)ζm+2],
ψ ≡hed − cos(θ1)i /ζ,
ζ ≡pd2 − 2d cos(θ1) + 1.
Appendix C: The stress tensor
21
,
(B10)
(B11)
(B12)
(B13)
(B14)
(B15)
Here we develop the expressions for the stress tensor in the cases discussed in the paper. The applied displacements
are symmetric about an axis passing through the centers of the two force dipoles. The expressions are taken from [41]
for the case of the displacement field given by Eqs. (5-6), excluding the first term in Eq. (5) which corresponds to
volume change. The stress tensor τ may be written in the following form in this case:
(C1)
τ =
∞Xn=0
τ (n)
rr
τ (n)
rθ
τ (n)
rϕ
τ (n)
θr
τ (n)
θθ
τ (n)
θϕ
τ (n)
ϕr
τ (n)
ϕθ
τ (n)
ϕϕ
,
Dn
eri
,
dθi
stress tensor τ = τ
G
R0
u0
, the elements of which are given by:
n+3 (n + 1)(n + 2)(cid:21) Yn (cosθi)
Cn
n+1 n(n2 + 3n − 2ν) +
Dn
eri
n+3 (n + 2)(cid:21) dYn (cosθi)
n+1 n(n2 − 2n − 1 + 2ν) −
The stress tensor is symmetriceτij =eτji and thus only six components are to be evaluated. We define the dimensionless
RR =2(cid:20)−
eτ (n)
eri
Rθ =2(cid:20) Cn
eτ (n)
n+1 (n2 − 2 + 2ν) −
eri
θθ =2(cid:26)(cid:20) Cn
eτ (n)
eri
ϕϕ =2(cid:26)(cid:20) Cn
eτ (n)
eri
Rϕ =eτ (n)
eτ (n)
n+3 (n + 1)2(cid:21) Yn(cosθi) −(cid:20) Cn
eri
n+3 (n + 1)(cid:21) Yn(cosθi) +(cid:20) Cn
eri
n+3(cid:21) dYn(cosθi)
eri
n+3(cid:21) dYn(cosθi)
eri
ctgθ(cid:27) ,
ctgθ(cid:27) ,
(C4)
n+1 n(n + 3 − 4nν − 2ν) −
n+1 (−n + 4 − 4ν) +
n+1 (−n + 4 − 4ν) +
Dn
Dn
eri
eri
dθi
dθi
Dn
Dn
(C2)
(C3)
(C5)
(C6)
θϕ = 0.
Appendix D: Derivation of Eq. (28)
22
Here we evaluate the integral (28) for the interaction energy of two "live" spherical force dipoles. Without loss of
generality, we take the initial isotropic displacement in the positive direction:
2 ·
1
2ZS
= − 2πu2
0
1
[(urFr) + (uθF θ)] 2πR2
0 sin θdθ(cid:12)(cid:12)(cid:12)(cid:12)er=1
2Z π
−→u · −→F ds = 2
er2 +eur11 +eur21(cid:19)(cid:20)−(cid:18) 4
11(cid:19)(cid:21) + (euθ11 +euθ21)(cid:16)−τ
−1(cid:26)(cid:18)−
0GR0Z 1
er3 + τ
1
22(cid:17)(cid:27) dcosθ(cid:12)(cid:12)(cid:12)(cid:12)er=1
Hereeur11, eur21, euθ11, euθ21 and τ are defined below [see Eqs. (12-15) and (C1-C6)]. The terms − 1
isotropic self-displacement created by each spherical force dipole and the corresponding force applied on it in response
er2 and − 4
er2 are the
(D1)
by the environment. Noting the fact that all the terms in (D1) are given in terms of Legendre polynomials Pn(cosθ)
and their derivatives dPn(cosθ)
dθ
and using their orthogonality, after integration and cancellation of the appropriate
terms we get:
0mCm + f Dr
0m (Dm −
√4πδm,0)i)
nmCm + f Dr
nm(Dm −
√4πδm,0)i)
16πu2
−
0GR0 1 + D0 −(−D0 +
∞Xm=0hf Cr
∞Xn=1(n(n + 3 − 4ν)Cn − (n + 1)Dn +
·(cid:2)n(n2 + 3n − 2ν)Cn − (n + 1)(n + 2)Dn(cid:3) ·
∞Xn=1((−n + 4 − 4ν)Cn + Dn +
∞Xm=0hf Cθ
2n + 1 !
· 2(cid:2)−(n2 − 2 + 2ν)Cn + (n + 2)Dn(cid:3) n(n + 1)
−
∞Xm=0hf Cr
2
4(2n + 1)
nmCm + f Dθ
nm(Dm −
√4πδm,0)i)
(D2)
In the FP cases the coefficients Cn, Dn for any n were either evaluated using the appropriate cancellation condition,
Eqs. (20-21) or assumed to be zero. In both FP cases the last two terms in Eq. (D1) vanish and the result becomes:
1
2ZS
2 ·
−→u · −→F ds = 16πu2
0GR0 1 + D0 −(−D0 +
∞Xm=0hf Cr
0mCm + f Dr
0m (Dm −
√4πδm,0)i)!
(D3)
In the VP cases on the other hand we do not calculate the coefficients C1 and D1 using the cancellation conditions,
but using the conditions (23-24) instead. Thus in VP cases neither the displacements, nor the stress vanish for n = 1.
Using (23-24) we get in these cases:
1
2ZS
2 ·
−→u · −→F ds = 16πu2
0GR0 1 + D0 −(−D0 +
∞Xm=0(cid:18) f Cθ
∞Xm=0(cid:18) f Cr
[(1 − 2ν)C1 + 3D1]
[(4 − 2ν)C1 − 6D1]
1
6
−
4
3
−
∞Xm=0hf Cr
0mCm + f Dr
0m (Dm −
1m − f Cr
1m
2
Cm +
f Dθ
1m − f Dr
1m
2
1m − f Cθ
1m
2
Cm +
1m − f Dθ
f Dr
1m
2
√4πδm,0)i)
Dm(cid:19)
Dm(cid:19)!
(D4)
The last two terms in Eq. (D4) are proportional to higher powers of Cn and Dn. It may be seen from Table I that
the coefficients Cn and Dn are inversely proportional to ed, and thus at large distances the last two terms in Eq. (D4)
may be neglected and Eq. (D4) reduces to Eq. (D3).
23
Appendix E: D0, C1, D1, C2, D2 and ∆E for arbitrary distance
In Table II we present the expressions for the coefficients D0, C1, D1, C2, D2 and for the dimensionless interaction
energy ∆eE without the large-distance assumption that appears in Table I.
Variable Size
Variable Size
Fixed Size
Fixed Size
Fixed Position
Variable Position
Fixed Position Variable Position
(F SF P )
(F SV P )
(V SF P )
(V SV P )
nmax
1
D0
C1
5(1−2ν) ed2
Λ1
− 15 ed4
2Λ1
2
λ1
Λ2
1
0
2
0
315(2−3ν) ed6
Λ2
− 15 ed4
2Λ3
63(2−3ν) ed3
Λ4
D1 −
ed[5 ed3
−(8−12ν)]
2Λ1
315(2−3ν)(7−8ν) ed6
Λ2
ed[5 ed3
−
−(8−12ν)]
2Λ3
63(2−3ν)(7−8ν) ed3
Λ4
C2
D2
∆ eE
−
0
0
10(1−2ν) ed2
Λ1
35[5(5−6ν) ed3 +(2−3ν)] ed7
Λ2
−
λ2
Λ2
λ3
Λ2
2
0
0
−
−
10(1−2ν) ed2
Λ3
−
35(5−6ν) ed7
Λ4
6(7 ed5
−28+40ν) ed2
Λ4
−
λ4
Λ4
∆ eE∞
2(1−2ν)
(5−6ν)
· 1
ed4
10(1−2ν)
(4−5ν)
· 1
ed6
−
2(1−2ν)
(5−6ν)
· 1
ed4
−
10(1−2ν)
(4−5ν)
· 1
ed6
TABLE II: Coefficients Cn and Dn of the first terms of the multipole expansion. nmax denotes the highest-order term taken
into account, ∆ eE ≡
Eint
E0
is the resultant dimensionless interaction energy and ∆ eE∞ is its asymptotic long-distance behavior
given also in Table I. ed = d
R0
is the dimensionless distance between the force dipoles.
We use the following expressions:
Λ1 = 5(5 − 6ν)ed6 − 30(1 − ν)ed5 + 10ed3 − (5 − 10ν)ed2 − (8 − 12ν),
Λ2 = 140(cid:0)20 − 49ν + 30ν2(cid:1)ed13 − 28(cid:0)242 − 403ν + 135ν2(cid:1)ed10 + 2520(5 − 6ν)ed8 − 35(cid:0)86 − 283ν + 216ν2(cid:1)ed7
− 1260(cid:0)10 − 27ν + 18ν2(cid:1)ed5 + 70(cid:0)2 − 7ν + 6ν2(cid:1)ed4 − 720(cid:0)35 − 92ν + 60ν2(cid:1)ed3 − 36(cid:0)14 − 41ν + 30ν2(cid:1) ,
Λ3 = 5(5 − 6ν)ed6 − 30(1 − ν)ed5 + 10ed3 − (8 − 12ν),
(E1)
(E2)
(E3)
Λ4 = 4(5 − 6ν)h7(4 − 5ν)ed10 − 35(2 − ν)ed7 + 126ed5 − 36(7 − 10ν)i ,
λ1 = 70(1 − 2ν)h10(5 − 6ν)ed3 − (2 − 3ν)ied4,
24
(E4)
(E5)
(E6)
λ2 =h35ed8(5 − 6ν) + 7ed5(2 − 3ν) − 20ed3(cid:0)35 − 92ν + 60ν2(cid:1) −(cid:0)14 − 41ν + 30ν2(cid:1)i 6ed2,
λ3 = 70ed4h2800(5 − 6ν)2(cid:0)10ν2 − 13ν + 4(cid:1)ed16 + 420(cid:0)330ν4 + 1411ν3 − 3961ν2 + 2898ν − 640(cid:1)ed13
− 50400(5 − 6ν)2(2ν − 1)ed11 − 14(cid:0)40230ν4 − 81594ν3 + 50371ν2 − 6691ν − 1586(cid:1)ed10
− 630(cid:0)7074ν4 − 17937ν3 + 15963ν2 − 5636ν + 580(cid:1)ed8 − 105(cid:0)3396ν4 − 11879ν3 + 15041ν2 − 8232ν + 1652(cid:1)ed7
− 14400(5 − 6ν)2(cid:0)20ν2 − 24ν + 7(cid:1)ed6 − 504(2 − 3ν)2(cid:0)375ν2 − 643ν + 275(cid:1)ed5 − 140(cid:0)6ν2 − 7ν + 2(cid:1)2 ed4
− 360(cid:16)14940ν4 − 46008ν3 + 52895ν2 −26907ν + 5110)ed3 − 18(2 − 3ν)2(cid:0)790ν2 − 1263ν + 497(cid:1)i ,
λ4 = 28edh1400(5 − 6ν)2(cid:0)4 − 13ν + 10ν2(cid:1)ed13 − 70(cid:0)8000 − 39830ν + 71471ν2 − 54759ν3 + 15030ν4(cid:1)ed10
+ 25200(5 − 6ν)2(1 − 2ν)ed8 − 525(cid:0)250 − 1465ν + 3163ν2 − 2988ν3 + 1044ν4(cid:1)ed7
− 945(cid:0)500 − 2180ν + 3431ν2 − 2277ν3 + 522ν4(cid:1)ed5 + 1890(2 − 3ν)2(cid:0)25 − 54ν + 29ν2(cid:1)ed4
− 7200(5 − 6ν)2(cid:0)7 − 24ν + 20ν2(cid:1)ed3 − 567(2 − 3ν)2(cid:0)300 − 673ν + 377ν2(cid:1)ed2
+ 540(cid:0)1750 − 9255ν + 18281ν2 − 15984ν3 + 5220ν4(cid:1)i .
(E7)
(E8)
[1] G. Salbreux, G. Charras, and E. Paluch, Trends in Cell Biology 22, 536 (2012).
[2] J. P. Winer, S. Oake, and P. A. Janmey, PloS one 4, e6382 (2009).
[3] M. Ghibaudo, A. Saez, L. Trichet, A. Xayaphoummine, J. Browaeys, P. Silberzan, A. Buguinb, and B. Ladoux, Soft Matter
4, 1836 (2008).
[4] N. Nisenholz, K. Rajendran, Q. Dang, H. Chen, R. Kemkemer, R. Krishnanb, and A. Zemel, Soft Matter 10, 7234 (2014).
[5] V. B. Shenoy, H. Wang, and X. Wang, Interface Focus 6, 20150067 (2016).
[6] C. A. Reinhart-King, M. Dembo, and D. A. Hammer, Biophysical Journal 95, 6044 (2008).
[7] T. Korff and H. G. Augustin, Journal of Cell Science 112, 3249 (1999).
[8] Q. Shi, R. P. Ghosh, H. Engelke, C. H. Rycroft, L. Cassereau, J. A. Sethian, V. M. Weaver, and J. T. Liphardt, Proceedings
of the National Academy of Sciences 111, 658 (2014).
[9] J. Notbohm, A. Lesman, P. Rosakis, D. A. Tirrell, and G. Ravichandran, Journal of The Royal Society Interface 12,
20150320 (2015).
[10] C.-L. Guo, M. Ouyang, J.-Y. Yu, J. Maslov, A. Price, and C.-Y. Shen, Proceedings of the National Academy of Sciences
25
109, 5576 (2012).
[11] A. Lesman, J. Notbohm, D. A. Tirrell, and G. Ravichandran, The Journal of Cell Biology 205, 155 (2014).
[12] J. Solon, I. Levental, K. Sengupta, P. C. Georges, and P. A. Janmey, Biophysical Journal 93, 4453 (2007).
[13] S. Y. Tee, J. Fu, C. S. Chen, and P. A. Janmey, Biophysical Journal 100, L25 (2011).
[14] A. J. Engler, S. Sen, H. L. Sweeney, and D. E. Discher, Cell 126, 677 (2006).
[15] H. Mohammadi, P. A. Janmey, and C. A. McCulloch, Biomaterials 35, 1138 (2014).
[16] S. He, Y. Su, B. Ji, and H. Gao, Journal of the Mechanics and Physics of Solids 70, 116 (2014).
[17] W. S. Leong, C. Y. Tay, H. Yu, A. Li, S. C. Wu, D.-H. Duc, C. T. Lim, and L. P. Tan, Biochemical and Biophysical
Research Communications 401, 287 (2010).
[18] X. Xu and S. A. Safran, Physical Review E 92, 032728 (2015).
[19] P. Rosakis, J. Notbohm, and G. Ravichandran, Journal of the Mechanics and Physics of Solids 85, 16 (2015).
[20] M. S. Rudnicki, H. A. Cirka, M. Aghvami, E. A. Sander, Q. Wen, and K. L. Billiar, Biophysical Journal 105, 11 (2013).
[21] X. Ma, M. E. Schickel, M. D. Stevenson, A. L. Sarang-Sieminski, K. J. Gooch, S. N. Ghadiali, and R. T. Hart, Biophysical
Journal 104, 1410 (2013).
[22] A. Abhilash, B. M. Baker, B. Trappmann, C. S. Chen, and V. B. Shenoy, Biophysical Journal 107, 1829 (2014).
[23] P. Ronceray, C. Broedersz, and M. Lenz, Proceedings of the National Academy of Sciences 113, 2827 (2016).
[24] H. Wang, A. S. Abhilash, C. S. Chen, R. G. Wells, and V. B. Shenoy, Biophysical Journal 107, 2592 (2014).
[25] Y. Shokef and S. A. Safran, Physical Review Letters 108, 178103 (2012).
[26] Y. Shokef and S. A. Safran, Physical Review Letters 109, 169901 (2012).
[27] G.Sines and R.Kikuchi, Acta Metallurgia 6, 500 (1958).
[28] T. Mura, Micromechanics of Defects in Solids (Cluwer Academic Publishers, 1991), 2nd ed.
[29] R. De, A. Zemel, and S. A. Safran, Biophysical Journal 94, L29 (2008).
[30] D. Ben-Yaakov, R. Golkov, Y. Shokef, and S. A. Safran, Soft Matter 11, 1412 (2015).
[31] U.S.Schwarz and S.A.Safran, Physical Review Letters 88, 048102 (2002).
[32] I. B. Bischofs, S. A. Safran, and U. S. Schwarz, Phys. Rev. E 69, 021911 (2004).
[33] U. S. Schwarz and S. A. Safran, Reviews of Modern Physics 85, 1327 (2013).
[34] K. M. Hakkinen, J. S. Harunaga, A. D. Doyle, and K. M. Yamada, Tissue Engineering Part A 17, 713 (2011).
[35] S. J. Mousavi and M. H. Doweidar, Physical Biology 11, 046005 (2014).
[36] J. Eshelby, Proceedings of the Royal Society of London A 241, 376 (1957).
26
[37] J. Eshelby, Proceedings of the Royal Society of London A 252, 561 (1959).
[38] R. Simes, Physica Status Solidi 30, 645 (1968).
[39] F. Bitter, Physical Review 37, 1527 (1931).
[40] J. Eshelby, Acta Metallurgica 3, 487 (1955).
[41] A. Lurie, Three-Dimensional Problems of the Theory of Elasticity (Interscience Publishers, 1964), english ed., chapter 6.
[42] G. B. Arfken, H. J. Weber, and F. E. Harris, Mathematical Methods for Physicists (Elsevier, 2013), 7th ed.
[43] J. D. Eshelby, Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering
Sciences 244, 87 (1951).
[44] J. N.Israelachvili, Intermolecular and Surface Forces (Harcourt Brace and Company, 1992), 2nd ed.
[45] J. Notbohm, A. Lesman, D. A. Tirrell, and G. Ravichandran, Integrative Biology 7, 1186 (2015).
[46] L. D. Landau and E. M. Lifshitz, Theory of Elasticity (Pergamon Press, 1970), 2nd ed.
[47] http://functions.wolfram.com/Polynomials/SphericalHarmonicY/20/01/01/0002/.
|
1701.01225 | 1 | 1701 | 2017-01-05T06:52:55 | Contact enhancement of locomotion in spreading cell colonies | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.CB"
] | The dispersal of cells from an initially constrained location is a crucial aspect of many physiological phenomena ranging from morphogenesis to tumour spreading. In such processes, the way cell-cell interactions impact the motion of single cells, and in turn the collective dynamics, remains unclear. Here, the spreading of micro-patterned colonies of non-cohesive cells is fully characterized from the complete set of individual trajectories. It shows that contact interactions, chemically mediated interactions and cell proliferation each dominates the dispersal process on different time scales. From data analysis and simulation of an active particle model, we demonstrate that contact interactions act to speed up the early population spreading by promoting individual cells to a state of higher persistence, which constitutes an as-yet unreported contact enhancement of locomotion. Our findings suggest that the current modeling paradigm of memoryless interacting active particles may need to be extended to account for the possibility of internal states and history-dependent behaviour of motile cells. | physics.bio-ph | physics | Contact enhancement of locomotion in
spreading cell colonies
Joseph d'Alessandro∗†‡, Alexandre Solon§, Yoshinori Hayakawa¶, Christophe Anjard†,
Fran¸cois Detcheverry†, Jean-Paul Rieu† and Charlotte Rivi`ere∗†
The dispersal of cells from an initially constrained location is a crucial aspect of many
physiological phenomena ranging from morphogenesis to tumour spreading. In such
processes, the way cell-cell interactions impact the motion of single cells, and in turn
the collective dynamics, remains unclear. Here, the spreading of micro-patterned
colonies of non-cohesive cells is fully characterized from the complete set of individual
trajectories. It shows that contact interactions, chemically mediated interactions and
cell proliferation each dominates the dispersal process on different time scales. From
data analysis and simulation of an active particle model, we demonstrate that contact
interactions act to speed up the early population spreading by promoting individual
cells to a state of higher persistence, which constitutes an as-yet unreported contact
enhancement of locomotion. Our findings suggest that the current modeling paradigm
of memoryless interacting active particles may need to be extended to account for the
possibility of internal states and history-dependent behaviour of motile cells.
Understanding how cell assemblies regulate
their motility is a major challenge of current bio-
physics.1 Indeed, collective effects in the motion
of cells play a crucial role in vivo in processes
such as wound healing2, tumour progression3 or
morphogenesis4. In understanding the often in-
tricate relationship between the behaviours at
the cellular level and the population scale, two
basic questions arise: how do cell-cell interac-
tions alter the properties of individual cell mo-
tion? How do they impact the population dy-
namics?
in descriptions based on the Fisher-Kolmogorov-
Petrovski-Piskunov (FKPP) equation7,8. How-
ever, the assumption of non-interacting cells
is often unwarranted9,10, as several types of
cell-cell interactions affecting the collective dy-
namics have been uncovered experimentally. A
first class involves long-range interactions, which
may be mediated by a chemical11,12 as in quo-
rum sensing, or by the substrate13. A second
class includes short-range contact interactions:
volume exclusion, cell-cell adhesion or contact
inhibition of locomotion (CIL)14, which acts to
change the direction of motion of a cell upon
contact with another cell.
7
1
0
2
n
a
J
5
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
5
2
2
1
0
.
1
0
7
1
:
v
i
X
r
a
The trajectory of a cell crawling on a surface is
akin to a correlated random walk characterized
by a persistence time beyond which the motion
becomes diffusive5,6. In the absence of interac-
tions, this would lead on long time to simple dif-
fusion dynamics at the colony level, as captured
∗<To whom correspondence should be addressed. E-
mail: joseph.d'[email protected], charlotte.riviere@univ-
lyon1.fr>
†Institut Lumi`ere Mati`ere, CNRS UMR 5306, Univer-
sit´e Claude Bernard Lyon 1, Universit´e de Lyon, Lyon,
69622 Villeurbanne Cedex, France
‡Current adress:
Institut Jacques Monod (IJM),
CNRS UMR 7592 et Universit´e Paris Diderot, 75013
Paris, France
§Department of Physics, Massachusetts Institute of
Technology, Cambridge, MA 02139, USA
¶Center for Information Technology in Education, To-
hoku University, Sendai 980-8578, Japan
Despite their local nature, contact interactions
have proven essential to the collective behaviour
of cells, at least for dense assemblies with den-
sity near close-packing. On the edge of a dense
colony, CIL15 or excluded volume16 combined
with a density gradient acts to bias the mo-
tion towards free space10,17,18, hence facilitating
the spreading of the colony9,16,19. This effect
is further reinforced by the tension created by
leader cells through adherens junctions17,20. In
the bulk of a tissue, force transmission through
adherens junctions21,22 (but also nematic align-
ment23,24 or simple volume exclusion25) can
lead to coordinated motion over several cell
sizes and induce active jamming and glassy be-
haviour26,27,28. The slowing down of tissue dy-
1
namics is especially clear in cell systems domi-
nated by CIL15, which reduces the cell persis-
tence29 or effective speed30.
Here, we investigate collective cell migration at
moderate density in assemblies lacking cell-cell
adhesion. In contrast to the high-density regime,
this region has received comparatively much less
attention so far. By studying the spreading
dynamics of micropatterned Dictyostelium dis-
coideum cell colonies, we find that cell-cell con-
tacts enhance the cell persistence, an effect that
we refer to as contact enhancement of locomo-
tion (CEL). This phenomenon results in a speed-
up of the colony spreading upon increasing the
packing fraction and defines a novel kind of
interaction, which, instead of acting instanta-
neously as a physical force, modifies the internal
state of the moving agents and their subsequent
behaviour.
A highly controlled model of cell
colony
We used vegetative Dictyostelium discoideum
(D.d.)
cells, which are often considered as a
benchmark for the amoeboid motility of fast-
moving cells31,32,33. Moreover, they are espe-
cially adapted to study the role of interactions in
the absence of strong cell-cell adhesion, as they
do not form such adhesions in nutrient-rich con-
ditions34. To experimentally mimic the disper-
sal of cells from an initial location in a repro-
ducible way, we constrained a controlled num-
ber of cells in a disk of diameter 320 µm, using
PDMS micro-stencils35,17 (Fig. 1a). Taking off
the micro-stencil, we let them migrate freely out-
wards and image the colony for durations rang-
ing from 8 h to 48 h (see snapshots in Fig. 1b
and Supplementary Movie 1).
We caracterize the colony spreading both at the
population and individual cell levels (Fig. 1c).
Making use of the circular symmetry, field quan-
tities such as the density (Fig. 1c) are aver-
aged over concentric rings and depend only on
time t and the distance r from the centre of the
colony. At short times, we first observe a de-
crease of the density in the centre of the colony
as it spreads to invade free space. Then, on
time scales of the order of the doubling time
β−1 ∼ 9 h, the density starts increasing uni-
formly because of cell divisions. Finally, after
2
about 40h the density saturates at a carrying
capacity ρmax ≈ 5 − 8 × 105 cell/cm2 (Fig. 1c).
The spreading of the colony is found to be faster
in the first few hours of the experiment. Us-
ing the single cell trajectories obtained by au-
tomated cell tracking, this observation can be
related to the average cell speed (see Supplemen-
tary Figure 1): After an initial increase, the cell
speed decreases until it reaches a low-motility
plateau at t ≈ 10 h. This behaviour at long
times is well explained by the overall regula-
tion of the motility through a secreted quorum-
sensing factor, which has been evidenced in our
group before11.
Indeed, when repeating the
spreading experiments with a continuous per-
fusion of fresh medium to rinse out secreted
molecules, the decrease in motility is suppressed
and the colony instead rapidly reaches a high-
motility plateau (see Supplementary Figure 1a).
As soon as the perfusion stops, the concentration
of quorum-sensing factors builds up and the cell
speed falls down.
Collective effects on the short-
time spreading
We now focus on the short-time spreading of
the colony. It strikingly reveals that the higher
the cell density, the faster the colony spreads
(Fig. 2a-b and Supplementary Movies 2-3). This
collective effect is seen on the density profiles
or on the gyration radius, Rg = phr2i, which
quantifies the size of the colony (see Fig. 2c --
e). To better characterise this effect, we com-
pute the radial component vr of the velocity of
each cell, and average over the colony. We find
that this averaged hvri exhibits a positive peak
around 100 min in the colonies with higher ini-
tial density N0 = 246 ± 66 (Fig. 2f and Sup-
plementary Figure 1). This implies that, in this
time frame, cells move outward in average.
The existence of a non-vanishing radial velocity
is not surprising in itself since it is the analogue,
for self-propelled particles36, of an outward dif-
fusive flux. However, one expects the peak to
be located at a time of the order of the persis-
tence time of the particles (see Fig. 4), which is
found around τp ∼ 5 min for D.d. cells in simi-
lar conditions11,37. On the contrary, the radial
velocity peak happens here on a much longer
time scale ∼ 100 min. It strongly suggests that
Figure 1: A highly-controlled experimental set-up gives full access to colony spreading dynamics at
both individual and population scales.
(a) Cartoon of the patterning technique. The cells are first deposited in a home-made well (brown)
on top of a PDMS stencil (light gray) made by soft-lithography techniques (top). After 45 min of
adhesion, the well and the stencil are removed (centre), creating an initial circular pattern of 320 µm
in diameter (bottom), whose spreading is then followed by time-lapse microscopy. (b) Snapshots of
a colony with N0 = 245 cells initially at t = 0 min (top) and t = 150 min (bottom). Scale bars:
200 µm. (c) Top: Evolution of the density profiles ρ(r, t) over 60 h (from blue to red) for one colony
with initially N0 = 349 cells. All the curves are separated by a 2 h interval. The first three curves
are drawn thicker to highlight the fast initial spreading of the colony. Bottom: Cell trajectories at
the edge of the initial spot, from t = 0 min to t = 60 min. Scale bar 100 µm.
an unknown effect speeds up the spreading on
this longer time scale. Importantly, this effect
neither originates from cell division (it happens
on a time-scale much shorter than the doubling
time) nor from distant chemically-mediated in-
teractions arising from bulk soluble molecules
(see Supplementary Figure 1) or from deposited
trails (see Supplementary Figure 2).
Most interestingly, we find that the amplitude
of the peak in radial velocity strongly depends
on the number of cells in the colony (Fig. 2f).
Thus, the spreading rate is collectively increased
through local interactions likely occurring when
cells are in contact. To understand this density-
dependent spreading dynamics, we now turn to a
more refined analysis of the motion of individual
cells.
3
Cell-cell contacts increase their
persistence
From our dataset of trajectories, we find that the
short-time speed-up of colony spreading is con-
trolled by a transient increase in the persistence
of the cells, the effect being more pronounced
the denser the initial colony. This statement, il-
lustrated by sample trajectories in Figure 3a-b
and Supplementary Figure 4, is motivated by a
body of quantitative measurements.
First, the cells appear more elongated, hence
more polarised, in denser colonies (Fig. 3a). It
is quantified by computing the cell contours' ec-
centricity, which increases with density at early
times before relaxing to values corresponding to
more isotropic cell shapes (see Supplementary
Figure 3).
Second, we used the coefficient of movement effi-
ciency (CME, see Methods) to estimate the per-
t=0 mint=150 min02004006000246x 105r (µm)ρ (cm−2)Time increasesabcFigure 2: Density-dependent colony spreading.
The experiments are divided in three groups to study the effect of the initial cell number N0. (a-b)
"Snapshots" of the colonies for two different groups N0 = 35 ± 12 (blue, left, 208 cells in total) and
246 ± 66 (red, right, 1229 cells in total), at t = 0 and t = 150 min. The positions of all cells in
each group of experiments are represented as coloured points. The dashed circle denotes the edge of
the stencil. (c-d) Normalised density profiles for each group at t = 0 min (c) and t = 150 min (d).
(e) Gyration radius Rg =phr2i of the colonies as a function of time. (f) Radial velocity hvri as a
function of time averaged over the colony (same colour code for every panel). The error bars are
the standard deviation (n = 6, 8, 5 experiments respectively for N0 = 35 ± 12, N0 = 97 ± 25, N0 =
246 ± 66).
sistence of the trajectories with good time and
space accuracy. This quantity, for a given in-
terrogation time ∆t, ranges from 0 for a mo-
tion with persistence time much smaller than
∆t to 1 for ballistic motion. From the spatio-
temporal evolution of the CME measured with
∆t = 5 min, it is clear that the persistence in-
creases with density (Fig. 3c). This is especially
pronounced at short times and near the periph-
ery of the colony, where the radial velocity map
also exhibits high values (Fig. 3d).
Finally, another quantitative characterisation is
provided by the velocity direction autocorrela-
tion function C(t0 − t) = hu(t0)· u(t)i, where
u is the direction of motion of a cell. The
simplest models of persistent motion (Ornstein-
Uhlenbeck, active Brownian particle or run-and-
tumble motion) all lead to an autocorrelation
function which decays exponentially over the
persistence time. In contrast, our data is better
described by a sum of two exponentials (Fig. 3e):
(1)
This form could equally arise because of two dis-
tinct populations of cells with different persis-
C(t) = c e−γt + c0 e−γ0t.
tence time or because of a bimodal motion of
individual cells41,44,42. However, the first possi-
bility can be ruled out since the relative weights
c and c0 are not constant in time.
Thus, all experimental clues point to the fact
that each cell
is able to increase its persis-
tence upon collision. For modeling simplicity,
although we cannot rule out completely a con-
tinuous change, we consider cells that switch be-
tween two modes of motion, a mode 1 whith per-
−1 and average duration τ1,
sistence time is Dr1
and a mode 2 with Dr2 < Dr1 and τ2. We used
this model to fit all experimental data, as ex-
plained in the Supplementary Information, as-
suming for simplicity that the mode of higher
persistence is ballistic (Dr2 = 0). Fig. 3e shows
representative examples for this fitting proce-
dure. As an output, we obtain the estimates
−1 =2 min and τ2 =10 min (see Supple-
Dr1
mentary Figure 5), and the fraction φ2(t) =
τ2/ (τ1 + τ2) of cells in mode 2 (Fig. 3f). Like
the radial velocity, φ2(t) reaches a maximum at
around 100 min, and exhibits an overall increase
with density, showing that higher densities pro-
4
Figure 3: Local interactions between cells lead to an increase in cell persistence.
(a -- d) Comparison of trajectory properties at low (hN0i = 35, top row) and high (hN0i = 246, bottom
row) initial densities. (a) Typical cell trajectory (plotted over 100 min), with "Start" denoting the
initial cell position, the white circle representing the border of the initial pattern and the arrows
pointing to collision events; (b) 64 examples of trajectories from t = 100 min to t = 130 min, picked
randomly over all cell trajectories; (c) spatio temporal dynamics of the CME with ∆t = 5 min and
(d) of the radial velocity vr (the dashed lines represent the border of the initial colonies). (e) Velocity
direction autocorrelation function at t = 67 − 133 min for various hN0i (symbols, same colour code
for panels e and f) and fits using the expression detailed in Supplementary Information (solid lines)
with D−1
r1 = 2 min and τ2 = 10 min. The thick black line represents the best single-exponential
('Single exp') fit for hN0i = 246, which misses the experimental behaviour at both very short and
long times. (f) Proportion φ2 of cells in mode 2 as extracted from the fit of the correlation functions
(the lines are guides for the eye). The error bars represent the 95% confidence interval for the fit
parameter. (g -- h) Normalised speed (g) and CME computed with ∆t = 5 min (h) for cells undergoing
a single collision at t = tcol within a 30 min interval. v (resp. CME), denote the basal speed (resp.
CME) before collision. The error bars show the SEM for the n = 464 pieces of trajectory from 232
collisions.
mote switching to the persistent mode.
Because this enhancement of persistence is not
chemically mediated (see Supplementary Figs. 1
and 2) and depends strongly on the density
(Figs. 2 and 3), we conclude that contacts --
understood here as collisions or short-range in-
teractions -- are the primary cause for the phe-
nomenon. Paralleling the CIL acronym, here-
after we refer to this effect as CEL, or contact
enhancement of locomotion. To confirm more
5
aCbhN0i = 35hN0i = 246bdd6004002000r (µm)012vr (µm.min-1)0200400t (min)6004002000r (µm)cc00.51CME (5 min)0200400t (min)01020t (min)0.10.20.30.4C(t)ehN0i = 35hN0i = 97hN0i = 246Single exp0200400t (min)00.10.20.30.4φ2f-10010t-tcol (min)11.1CME/CMEh11.21.4v/vgdirectly the existence of this phenomenon, we
look at the statistics of cell-cell contacts. As de-
scribed in the Supplementary Information, we
retain only "clean" contacts (between cells un-
dergoing no other collision for 15 min before and
after). We compare the speed and CME of each
single cell before and after collision and then
average over all available data (see Fig. 3g -- h,
and Suplementary Figures 10 -- 12). Both com-
puted quantitites exhibit a significant transient
increase, demonstrating that CEL is indeed re-
sponsible for the density-dependent spreading
rate of the colonies. Note that the analysis of
the angular scattering (see Supplementary Fig-
ures 8-9) show that the cell-cell contacts have no
aligning effect likely to promote collective mo-
tion and to increase the spreading rate by itself.
Comparison with minimal active
particle models
To support our experimental findings, we now
investigate several minimal active particle mod-
els. Discarding other types of contact interac-
tions, we show that a collision-induced increase
in persistence is necessary and sufficient to ac-
count for the salient features of the short-time
dynamics of the colony spreading.
Let us consider self-propelled hard disks mov-
ing at a constant velocity v.
In addition, the
direction of motion θ of a particle is subject to
rotational diffusion with coefficient Dr. The mo-
tion of the ith particle is thus governed by the
following equations
fij(ri − rj)
(2)
∂tri = vu(θi) +X
∂tθi =p2Drηi(t),
j6=i
(3)
where u(θi) = (cos θi, sin θi) and ηi is a delta-
correlated Gaussian white noise with zero mean
and unit variance. fij is the steric repelling force
exerted by particle j on particle i (see Methods
for more details). As such, Eqs. (2-3) describe
the active Brownian particle (ABP) model, well-
studied as a minimal model of phase-separating
active particles30,38, and a suitable basis for the
modelling of the persistent random motion of
cells. Consistently with the experiments, we
take v = 5 µm.min−1 and initially place the par-
ticles in a disk of 320 µm in diameter. The re-
sulting average radial velocity and radius of gy-
ration measured in the simulations are shown as
a function of time in Fig. 4.
Indeed,
First, we checked whether this simple interac-
tion rule could yield a density-dependent spread-
ing.
if we think of moving cells as
hard-core spheres undergoing a persistent ran-
dom walk, the excluded-volume (EV) between
the cells gives rise to an outward pressure38.
This effect is also present for Brownian hard
spheres, for which it can be taken into account
by an effective diffusion coefficient increasing
with concentration39. However, in the present
experiments where cells are relatively sparse --
with packing fractions up to 0.3 -- this pressure
is not expected to play an important role and in-
deed, simulations of Eqs. (2)-(3) with only hard-
core repulsion exhibit no effect of density (Fig. 4,
left).
a contact: ΓP
We then further implemented the effect of CIL
that is, upon collision, cells actively reorient
away from the contact. To that end, we added
an angular repulsion to the equations of motion,
in the form of a torque acting on cells undergoing
j6=i H(krj − rik − σr) sin(θi − βij)
in equation (3), where βij = arg(rj − ri), H is
the Heaviside step function implementing the fi-
nite radius of interaction and σr is the contact
distance (see Methods). Qualitatively, one can
imagine that by reorienting the direction of mo-
tion of particles toward free space, active reori-
entation could explain the experimental data15.
However, even with a large value Γ = 100 min−1
corresponding to quasi-instantaneous reorienta-
tion, the density-dependent increase in average
radial velocity and spreading rate is an order
of magnitude smaller than in the experiment
(Fig. 4, centre). In addition, the (very limited)
peak in radial velocity appears at a very early
time, at odds with the experimental observation.
In both situations, the rotational diffusion was
set to D−1
r = 5 min to match the average per-
sistence time of experimental trajectories.
Finally, guided by the experimental observa-
tions, we tested a minimal model of CEL: After
each contact, two particles involved in a collision
enter a mode of high persistence (Fig. 4, right).
For simplicity, the speed is kept constant and the
motion is taken to be ballistic in this high per-
sistence (mode 2) state. The particles then relax
to the basal (mode 1) state at a rate λ2 = τ−1
2 .
We fix D−1
r1 = 2 min and τ2 = 10 min, as deter-
mined from the fit of the experimental data (see
6
Figure 4: Spreading colonies in a particle-based model mimicking the experiments.
Radial velocity hvri (dashed lines, left axes) and colony radius Rg (solid lines, right axes) in simulated
colonies with different numbers of particles N0 in the three models considered. The cartoons on the
top row illustrate the different contact rules: Excluded-volume only (EV, left), excluded-volume
and angular repulsion (EV + CIL, center), excluded-volume and contact enhancement of locomotion
(EV + CEL, right).
Fig. 3). The proportion φ2 of cells in the persis-
tent mode (shown in Supplementary Figure 6a)
evolves with time in a density-dependent man-
ner due to changes in the collision frequency.
We find that this model captures well the collec-
tive spreading of the colony. Indeed, as shown
in Fig. 4, the amplitude and density-dependence
of the peak in radial velocity, as well as the
faster increase of Rg at higher density, are well
captured. Overall, given the simplicity of the
model, the agreement between simulations and
experiments appears surprisingly good (see Sup-
plementary Movies 2 -- 5 for a visual comparison).
Discussion
We studied the dynamics of spreading colonies of
D.d. cells. Using a micro-fabrication technique,
we were able to produce initial colonies with con-
trolled shape and number of cells. We showed
that the long-time dynamics of the spreading is
controlled by cell divisions and long-range in-
teractions through a quorum-sensing factor. On
the contrary, these two effects are absent from
the short-time dynamics, thus allowing us to
study the effect of cell-cell contacts. We found
that cell contacts enhance the spreading of the
colony by increasing the speed and the persis-
tence time of the cells motion. This contact en-
hancement of locomotion is further supported by
a simple active particle model reproducing the
main characteristics of the experimental data.
The phenomenon of "CEL" that we have de-
scribed here seems a priori different from CIL,
which acts to change the direction of motion
of colliding cells. However,
they are not mu-
tually exclusive (see Supplementary Figure 12)
and could even share a common microscopic ori-
gin.Indeed, the current explanation for CIL is
that the protrusions driving the motion are in-
hibited in the contact region15,45. Other protru-
sions can thus develop elsewhere on the cell's
periphery,
leading to a new direction of mo-
tion. We could hypothesize that, similarly, ei-
ther the inhibition of ruffling in the contact re-
gion or the stabilisation of the new protrusions
reinforces the new polarity thereby increasing
the speed and persistence of the motion. An
important difference between the two effects is
that CEL involves memory while CIL is usu-
ally modeled as an "instantaneous" contact pro-
cess46,47,48,49. The two could have different rel-
ative importances depending on the situation.
For example, in dense cell clusters, the reorien-
tation induced by CIL forces the particles on the
7
0200400100200300400t (min)Rg (µm)020040000.511.5t (min)〈vr〉 (µm.min−1) 0200400t (min) EVEV+CILEV+CELN0 = 20N0 = 50N0 = 100N0 = 200N0 = 300edge to move outward4,47 whereas, as we saw, in
sparser colonies the increase in persistence is the
dominating effect.
The importance of contact interactions in vivo is
not completely understood. CIL has been found
to play an important role in neural crest migra-
tion4 and the loss of heterotypic CIL (between
different cell types) is thought to be crucial in
the invasion of healthy tissues by cancer cells3,50.
Similarly, CEL could be an advantage in invad-
ing the surrounding environment efficiently, in
the case of a less cohesive group of cells. Such a
situation can be encountered during the escape
of highly metastatic and invasive cancer cells, as
well as in cell morphogenesis and microbial dis-
persal. More generally, it should be noticed that
CEL is reminiscent of escape mechanisms, found
in various organisms, which involve a temporary
change of the motile behaviour and can lead to
surprising collective effects51.
Finally, the new type of interaction uncovered
here opens questions for active matter. Indeed,
it exemplifies the wide range of possible interac-
tions between active particles, compared to the
usual "physicist's particles", which can lead to
a rich phenomenology. It opens the door to fur-
ther studies of interactions that act on an ad-
ditional internal degree of freedom, which could
exhibit other interesting effects.
METHODS
coated on the mould for 1 min at 750 rpm to a target
thickness of 70 µm. The squares were cut and peeled
off. Usually a thin PDMS membrane obstructed the
hole. It was then removed with a surgical blade un-
der the microscope at low magnification.
The square stencil was stuck on the ground of a
3.5 cm wide culture dish and a homemade small plas-
tic well was stuck on it using silicon seal. A droplet
of medium was deposited into the well and the sam-
ple was placed under vacuum for 15 min to help the
medium enter the central hole of the stencil and wet
the dish's surface.
The cell suspension was added in the well and the
sample was placed in the incubator for 45 min to let
the cells sediment and adhere. Then, the plastic well
and the stencil were removed with surgical tweez-
ers. Last, the spreading colony was imaged using a
slightly defocused bright-field microscope (TE2000,
Nikkon) at 10X magnification and a wide-field An-
dor Zyla sCMOS camera. A time-lapse movie was
recorded for up to 48 h using MicroManager software
with a 20 s time-interval, while the temperature was
kept constant at 22.5◦C.
For perfusion experiments, we designed a macroflu-
idic chamber by sealing the culture dish with an
adapted cover containing an input and an output
tube. The former was linked to a 1 L supply bottle
of fresh HL5 medium under controlled overpressure
(OB1 controller, Elveflow) while the latter was linked
to a disposal bottle. All the system was closed ster-
ilely. We used a flow rate of 100 mL/h so that the
chamber volume of about 10 mL was completely re-
newed every 6 minutes. We were thus able to main-
tain a stable medium renewal over 9 hours.
Cell culture
Image processing
We used Dictyostelium discoideum cells from the
strain AX2.
The cells were cultured on cell-
culture-treated Petri dishes (BD Falcon) in HL5
medium with glucose (Formedium) and kept in a
temperature-controlled incubator at 22.5◦C, with a
doubling time β−1 ∼ 9 h. Before every experiment,
the cells were detached from the dish, centrifuged
5 min at 663g, harvested and resuspended at the
seeding density.
Sample preparation
A reusable mould on Si wafer comprising an array
of squares with circular pillars of height ∼ 150 µm
and diameter 320 µm in the centre was fabricated in
SU8 photoresist using classical soft lithography tech-
niques and its surface was silanized to make it non-
adherent. PolyDiMethylSulfoxyde (PDMS, Corning)
mixed with curing agent at a 1:10 mass ratio was spin
based
minimization
ImageJ macros
trajectories were
on
the
Then the
The cells' positions were retrieved using home-
'Find
made
Maxima' built-in function.
indi-
reconstructed with a
vidual
squared-displacement
algorithm
(http://site.physics.georgetown.edu/matlab/)
and the data analysed using homemade Matlab
programs.
In particular, the CME was defined as:
2 ) − r(t − δt
2 )k
kv(t0)kdt0
CMEδt(t) =
kr(t + δt
R t+ δt
2
t0=t− δt
2
(4)
Simulations
Simulations were carried out by integrating the
Langevin equations Eqs. (2-3) using a Euler integra-
tion scheme with time steps ∆t = 10−3 min. The
8
hard-core repulsion between particles is modelled
by a Weeks-Chandler-Andersen potential V (r) =
(cid:1)6i + 1 if r < 21/6σ and 0 otherwise,
(cid:1)12 −(cid:0) σ
4h(cid:0) σ
r
r
where σ = 10 µm is the particle diameter. We define
two particles as being in contact when their relative
distance r < σr = 21/6σ.
In the simulations with
CIL, the torque term is turned on only during the
contacts, when r < σr. In the simulation including
CEL, a contact triggers a ballistic run which lasts for
an exponentially distributed time with rate λ2.
References
[1] Travis, J. Mysteries of the cell: Cell biol-
ogy's open cases. Science 334, 1051 (2011).
[2] Friedl, P. & Gilmour, D. Collective cell mi-
gration in morphogenesis, regeneration and
cancer. Nat Rev Mol Cell Biol 10, 445 -- 457
(2009).
[3] Friedl, P. & Wolf, K. Tumour-cell invasion
and migration: diversity and escape mech-
anisms. Nature Reviews Cancer 3, 362-374
(2003).
[4] Carmona-Fontaine, C. et al. Contact inhi-
bition of locomotion in vivo controls neu-
ral crest directional migration. Nature 456,
7224 (2008).
[5] Selmeczi, D. et al. Cell motility as random
motion: A review. European Physical Jour-
nal: Special Topics 157, 1 -- 15 (2008).
[6] Li, L., Norrelkke, S.F. & Cox, E.C. Per-
sistent cell motion in the absence of exter-
nal signals: A search strategy for eukaryotic
cells. PLoS ONE, 3, 5 (2008).
[7] Kolmogorov, A., Petrovskii, I. & Piscounov,
N. A study of the diffusion equation with
increase in the amount of substance, and its
application to a biological problem. Math.
Mech. 1, 1-âĂŞ25 (1937).
[8] Simpson, M.J. et al. Quantifying the roles
of cell motility and cell proliferation in a
circular barrier assay. Journal of the Royal
Society, Interface / the Royal Society 10,
20130007 (2013).
[9] Sengers, B.G., Please C.P. & Oreffo R.O.C.
Experimental characterization and compu-
tational modelling of two-dimensional cell
spreading for skeletal regeneration. Journal
9
of the Royal Society, Interface / the Royal
Society, 4, 1107 -- 1117 (2007).
[10] Marel, A.K. et al. Flow and Diffusion in
Channel-Guided Cell Migration. Biophys.
J., 107, 1054 -- 1064 (2014).
[11] Gol´e, L., Rivi`ere, C., Hayakawa, Y. & Rieu,
J.P. A quorum-sensing factor in vegetative
Dictyostelium Discoideum cells revealed by
quantitative migration analysis. PLoS ONE
6, 1 -- 9 (2011).
[12] Phillips, J. & Gomer, R. A secreted pro-
tein is an endogenous chemorepellant in
Dictyostelium discoideum. Proc. Natl Acad.
Sci. USA, 109, 10990 -- 10995 (2012).
[13] Angelini, T.E., Hannezo, E., Trepat, X.,
Fredberg, J.J. & Weitz, D.A. Cell migration
driven by cooperative substrate deforma-
tion patterns. Phys. Rev. Lett. 104, 168104
(2010).
[14] Abercrombie, M. & Heaysman, J.E. Ob-
servations on the social behaviour of cells
in tissue culture: I. Speed of movement of
chick heart fibroblasts in relation to their
mutual contacts. Experimental cell research
5, 111 -- 131 (1953).
[15] Mayor, R. & Carmona-Fontaine, C. Keep-
ing in touch with contact inhibition of loco-
motion. Trends in cell biology 20, 319 -- 328
(2010).
[16] Dyson, L. & Baker, R.E. The importance of
volume exclusion in modelling cellular mi-
gration. J. Math. Biol. 71, 679 -- 711 (2014).
[17] Serra-Picamal, X. et al. Mechanical waves
during tissue expansion. Nat. Phys. 8, 628 --
634 (2012).
[18] Nnetu, K.D., Knorr, M., Strehe, D., Zink,
M. & Kas, J.A. Directed persistent mo-
tion maintains sheet integrity during multi-
cellular spreading and migration. Soft Mat-
ter 8, 6913 (2012).
[19] Yates, C.A., Parker, A. & Baker, R.E.
Incorporating pushing in exclusion-process
models of cell migration. Phys. Rev. E 91,
052711 (2015).
[20] Sep´ulveda, N. et al. Collective cell motion
in an epithelial sheet can be quantitatively
described by a stochastic interacting parti-
cle model. PLoS Comput. Biol. 9, e1002944
(2013).
[32] Friedl, P. & Wolf, K. Plasticity of cell mi-
gration: a multiscale tuning model. J. Cell
Biol. 188, 11 -- 9 (2010).
[21] Petitjean, L. et al. Velocity fields in a col-
lectively migrating epithelium. Biophysical
J. 98, 1790 -- 1800 (2010).
[33] Levine, H. Learning Physics of Living Sys-
tems from Dictyostelium. Phys. Biol. 11,
053011 (2014).
[22] Tambe, D.T. et al. Collective cell guid-
ance by cooperative intercellular forces.
Nat. Mat. 10 469 -- 475 (2011).
[23] Coburn, L., Cerone, L., Torney, C., Couzin,
I.D. & Neufeld, Z. Interactions lead to co-
herent motion and enhanced chemotaxis
of migrating Cells. Phys. Biol. 10, 046002
(2013).
[24] Duclos, G., Garcia, S., Yevick, H.G. & Sil-
berzan, P. Perfect nematic order in con-
fined monolayers of spindle-shaped cells.
Soft Matter 10, 2346 (2014).
[25] Londono C. et al. Nonautonomous contact
guidance signaling during collective cell mi-
gration. Proc. Natl Acad. Sci. USA 111
1807 -- 1812 (2014).
[26] Angelini, T.E. et al. Glass-like dynamics of
collective cell migration. Proc. Natl Acad.
Sci. USA 108, 4714 -- 4719 (2011).
[27] Park, J.-A. et al. Unjamming and cell shape
in the asthmatic airway epithelium. Nat.
Mat. 14, 1040 -- 1049 (2015).
[28] Garcia, S. et al. Physics of active jam-
ming during collective cellular motion in a
monolayer. Proc. Natl Acad. Sci. USA 112,
15314 -- 15319 (2015).
[29] Vedel, S., Tay, S., Johnston, D.M., Bruus,
H. & Quake, S.R. Migration of Cells in a
Social Context. Proc. Natl Acad. Sci. USA
110, 129-134 (2013).
[30] Fily, Y. & Marchetti, M.C. Athermal
Phase Separation of Self-Propelled Parti-
cles with No Alignment. Phys. Rev. Lett.
108, 235702 (2012).
[31] Friedl, P., Borgmann, S. & Brocker, E. B.
Amoeboid leukocyte crawling through ex-
tracellular matrix:
lessons from the Dic-
tyostelium paradigm of cell movement.
Journal of Leukocyte Biology 70, 491-509
(2001).
[34] Coates, J. C. & Harwood, A. J. Cell-cell ad-
hesion and signal transduction during Dic-
tyostelium development. J. Cell Sci. 114,
4349-4358 (2001).
[35] Poujade, M. et al. Collective migration of
an epithelial monolayer in response to a
model wound. Proc. Natl Acad. Sci. USA
104, 15988 -- 15993 (2007).
[36] Cates, M.E. & Tailleur, J. When are
active Brownian particles and run-and-
tumble particles equivalent? Consequences
for Motility-induced Phase Separation. Eu-
rophys. Lett. 101, 20010 (2013).
[37] Bosgraaf, L. & Van Haastert, P.J.M. The
ordered extension of pseudopodia by amoe-
boid Cells in the absence of external cues.
PLoS ONE 4, 4 (2009).
[38] Solon, A.P. et al. Pressure and phase
equilibria in interacting active Brownian
spheres. Phys. Rev. Lett. 114, 198301
(2015).
[39] Bruna, M. & Chapman, S.J. Excluded-
volume effects in the diffusion of hard
spheres. Physical Review E 85, 011103
(2012).
[40] Peruani, F. & Morelli, L.G. Self-propelled
particles with fluctuating speed and direc-
tion of motion in two dimensions. Phys.
Rev. Lett. 99, 010602 (2007).
[41] Potdar, A.A., Jeon, J., Weaver, A.M.,
Quaranta, V. & Cummings, P.T. Hu-
man mammary epithelial cells exhibit a
bimodal correlated random walk pattern.
PLoS ONE 5, e9636 (2010).
[42] Metzner, C. et al. Superstatistical analy-
sis and modelling of heterogeneous random
walks. Nat. Commun. 6, 7516 (2015).
[43] Selmeczi, D. et al. Cell motility as persis-
tent random motion: theories from experi-
ments. Biophys. J. 89, 912 -- 931 (2005).
10
AUTHOR CONTRIBUTIONS
J.d.A., J.P.R. and C.R. designed experiments; J.d.A.
performed experiments and analysed experimental
data; J.d.A. and A.S. conceived the particle-based
models; A.S. performed simulations and analysed
simulation data; C.A. contributed to design of ex-
periments in Supplementary Figure 1 and provided
AprA− cells; F.D. computed the analytical results
on bimodal trajectories and helped with the fitting
procedure; Y.H. assisted in the data analysis and in-
terpretation; J.d.A. and A.S. wrote the manuscript;
F.D., J.P.R. and C.R. made substantial contribu-
tion to the manuscript; all authors discussed and
interpreted the data, read and commented on the
manuscript; J.P.R. and C.R. supervised the project.
[44] Maiuri, P. et al. Actin Flows Mediate a
Universal Coupling between Cell Speed and
Cell Persistence. Cell 161, 374 -- 386 (2015).
[45] Stramer, B.A. & Mayor, R. Mechanisms
and in vivo functions of contact inhibition
of locomotion. Nat. Rev. Mol. Cell. Biol.
118 (2016).
[46] Davis, J.R., et al. Emergence of embry-
onic pattern through contact inhibition of
locomotion. Development 139, 4555-4560
(2012)
[47] Zimmermann, J., Camley, B.A., Rappel,
W.-J. & Levine, H. Contact inhibition of
locomotion determines cell-cell and cell-
substrate forces in tissues. Proc. Natl Acad.
Sci. USA 113, 2660-2665 (2016).
[48] Camley, B.A., Zimmermann, J., Levine,
H. & Rappell, W.-J. Emergent collec-
tive chemotaxis without single-cell gradi-
ent sensing. Phys. Rev. Lett. 116, 098101
(2016).
[49] Szabo, A., et al. In vivo confinement pro-
motes collective migration of neural crest
cells. J. Cell. Biol. 213, 543 -- 555 (2016).
[50] Abercrombie, M. Contact inhibition and
malignancy. Nature 281, 259 -- 262 (1979).
[51] Ramdya, P. et al. Mechanosensory in-
in
teractions drive collective behaviour
Drosophila. Nature 519, 233 -- 236 (2015).
[52] Heid, P. J. et al. The role of myosin heavy
chain phosphorylation in Dictyostelium
motility, chemotaxis and F-actin localiza-
tion. J. Cell Sci. 117, 4819 -- 4835 (2004).
ACKNOWLEDGEMENTS
The authors are grateful to R. Fulcrand for his help
in micro-fabrication, to V. Hakim for stimulating dis-
cussions and to C. Cottin-Bizonne for her comments
on the manuscript. J.d.A. has been partially sup-
ported by the Fondation ARC pour la Recherche
sur le Cancer and by the Programme d'Avenir Lyon-
Saint ´Etienne. A.S. acknowledges funding through a
PLS fellowship of the Gordon and Betty Moor foun-
dation. J.d.A., C.R. and J.P.R. belong to the CNRS
consortium CellTiss and to the LIA ELyTLab.
11
Supplementary information
1 An effect of local interactions
Supplementary Figure 1: The cells' outward motion is not due to a large-scale chemical communi-
cation.
Colony-averaged speed (top row) and radial velocity (bottom row) in various conditions. (a) Medium
exchange. Control (blue) and samples with continuous medium exchange (red). The speed is main-
tained constant until the exchange is stopped (vertical dashed line), whereas the radial motion hvri
is robust to medium exchange. (b) aprA− cells still exhibit outward motion, although they do not
secrete AprA, the only known endogenous chemorepellent in Dictyostelium discoideum (c) Cells in
highly conditioned medium (HCM), prepared by letting cells in culture in it before filtering, so that
all secreted molecules are already at a high concentration. The cell motion is affected but there is
still a density-dependent peak in hvri(t).
We checked that long-distance interactions were not responsible for the enhancement of the colony
spreading rate. To that end, we measured the average speed hvi and radial velocity hvri during
the spreading of cell colonies in three different control conditions, as described below. We looked
especially at the presence of the peak in hvri(t), which, in our experiments, is the macroscopic
signature of the collective enhancement of the spreading. Firstly, we designed a fluidic system that
allowed to continuously change the sample's medium, so that any secreted (or depleted) molecule was
rinsed out (or rescued), hence preventing any large-scale chemical sensing such as chemorepulsion
or quorum-sensing. It efficiently suppressed the overall regulation of cell motility through a known
quorum-sensing factor (QSF)11 (Supplementary Figure 1a, top), but did not affect the outward
motion (Supplementary Figure 1a, bottom) which shows that the collective effect is still present.
Secondly, we used aprA− cells which do not produce the protein AprA, so far the only endogenous
chemorepellent molecule known for Dictyostelium discoideum 12. Although the motility of these cells
is slightly different from the wild-type cells, resulting in slightly modified colony dynamics, the main
effect of outward motion was still observed (Supplementary Figure 1b). Last, experiments were done
in highly conditioned medium (HCM). This medium is prepared by letting cells in culture in fresh
HL5 medium for typically 2 days, so that it is supplemented with molecules secreted by the cells,
at high concentration. Thus, one would expect the concentrations in slowly degraded molecules
to be above the saturation of their detection by the cells, hence screening any additional secretion
12
020040060080034567t (min)〈v〉 (µm.min−1)A a. Medium change020040060080000.511.52t (min)〈vr〉 (µm.min−1)02004000246t (min)〈v〉 (µm.min−1) b. AprA− cells020040000.20.40.60.8t (min)〈vr〉 (µm.min−1)0200400012345t (min)〈v〉 (µm.min−1) c. Highly conditionedmedium 020040000.20.40.6t (min)〈vr〉 (µm.min−1) 〈N0〉=35〈N0〉=74〈N0〉=292during the experiment. Although the dynamics is again modified mainly due to the presence of
QSF at high concentration in HCM11, the collective effect of outward motion is still apparent in
this situation (Supplementary Figure 1c).
We also tested the hypothesis of cell-cell communication through the deposition of chemicals on
the surface as follows. First, the sample dishes were treated by letting cells adhere on their entire
surface at high density for 45 min. Then the cell layer was washed out and fresh cells were added
at a low, homogeneous density with fresh medium for imaging and the single cells were tracked
after 45 min adhesion. The controls include prior incubation with FM or HCM but no cell layer. If
cell-cell communication through deposited trails on the surface affects the motility, there should be
measurable differences between those conditions. Conversely, since the system is homogeneous and
isotropic and the cells undergo almost no cell-cell contact, no effect of geometry or interaction is
expected to interfere with those prior treatments. Yet, the probability density functions (PDFs) of
speed and CME in the three conditions are well overlaid, showing no effect of a putative chemical
deposition mechanism. Those PDFs compare very well with those measured in the low density
colonies, while they differ from the high density colonies, especially around the peak of hvri(t) at
t = 1 -- 2 h (Supp. Fig. 2).
Supplementary Figure 2: Modifications of the surface by the cells do not affect cell motility.
Distribution of speed (left) and CME with ∆t = 5 min (right) computed from trajectories with
δt = 1 min, gathered for all times t = 0 − 1 h (top) and t = 1 − 2 h (bottom) with various surface
treatments (solid lines) or in spreading colonies (dashed lines). Cells ⇒ FM: cells were seeded at
high density, let adhere for 45 min and washed out before adding fresh cells in fresh medium (FM)
for imaging. FM ⇒ FM: the sample dish was filled with FM for 45 min before cells and FM were
added for imaging. HCM ⇒ FM: the sample dish was filled with HCM for 45 min before cells and
FM were added for imaging.
Taken together, these results show that a large-scale communication using secreted or deposited
molecules is very unlikely to be at the origin of the density-dependent spreading at short times.
In consequence, the interactions behind this effect must be local, whether mediated by actual cell-
cell contacts or by rapidly degraded, locally accumulated chemicals. Although the latter is not
inconceivable, it is less likely at stake and can be described as an "effective contact" interaction.
13
00.10.20.3PDF (A.U.)hN0i=35hN0i=24600.10.20.3PDF (A.U.)t=0 -- 1hCells => FMFM => FMHCM => FM0510152025hvi (µm.min-1)00.10.20.30.4PDF (A.U.)t=1 -- 2h00.51CME00.10.20.30.4PDF (A.U.)2
Increase in single cell persistence
Supplementary Figure 3: Cells tend to be more elongated at high density.
Time evolution of the average eccentricity of the cell shapes' obtained by fitting ellipses for different
initial densities hN0i. The error bars are the standard deviations of the e distributions.
q
In Dictyostelium discoideum, the more persistent cells are also the more elongated ones, while the
ones that exhibit rounded shapes usually move in a more random fashion11,52. A way to measure the
elongation of cells is to fit their shape with ellipses and compute the eccentricity e =
a, where
a and b are respectively the long and short axes of the ellipse. The eccentricity varies between 0 for
a circle and 1 for an infinitely elongated ellipse. For instance, e = 0.5 corresponds to a/b = 1.33,
e = 0.7 to a/b = 1.96 and e = 0.9 to a/b = 5.26. We measured the colony average of e as a function
of time for different cell densities (Supplementary Figure 3). At early times, hei goes up with cell
density. It then relaxes to hei ≈ 0.55 for all hN0i. Although this elongation seems to be partly
defined prior to the release of the stencil (see hei at t = 0), a peak still appears in the densest
condition, concomitantly with the peaks in hvri and φ2.
As for the persistence of cell trajectories, it increases transiently with no detectable pre-set depen-
dence on density. To illustrate this fact, we show randomly sampled trajectories -- all of the same
30 min duration -- taken from experiments with low (hN0i = 35) or high (hN0i = 246) cell densities
at different times (Supplementary Figure 4). At early times, most trajectories show a low persis-
tence for both densities. This changes in the high density case where trajectories appear "unfolded"
around t = 100 min, demonstrating the existence of runs with persistence time comparable to the
30 min path duration. At later times, the motion becomes again less persistent.
1 − b
14
Supplementary Figure 4: Increase in single cell persistence.
Example of single cell trajectories from the lowest (hN0i = 35) or highest (hN0i = 246) cell density
experiments, at times t = 0− 30 min, t = 100− 130 min, t = 200− 230 min, t = 300− 330 min. The
persistence is seen to increase transiently around t = 100 min in the high density case, while at low
density it does not change markedly in time. Grid spacing: 200 µm.
15
Time0 - 30 min100 - 130 min200 - 230 min300 - 330 minhN0i = 35hN0i = 2463 Bimodal persistent motion
We found two modes in the decay of the velocity direction auto-correlation function (VDACF). As
mentioned, either those two modes arise from two separate populations of cells, each having a single
correlation time, or every single cell exhibit both modes. The first hypothesis is ruled out since it
does not allow changes in the weights of the modes. In the second hypothesis, the motion of each
cell follows a process with two characteristic times: this is the case of bimodal motion41,44,42, but
other models have this same property43. Here we focus on the first option because it offers the
simplest interpretation of a change in the weights c and c0 in Equation (1). In the model proposed
by Selmeczi et al.43, it would mainly involve tuning their α parameter, which is the strength of a
memory in a modified O.U. process: the underlying principle -- regulation of the relative importance
of two time-scales -- is the same, but the formulation involves a higher degree of complexity which
seems unnecessary here.
Model. We first characterise analytically a simple model of bimodal persistent motion. Assume
a particle moves in the plane with velocity of constant magnitude v0. Its orientation is subject to
rotational diffusion, but with a coefficient that alternates between two values Dr1 and Dr2. The
times spent in mode 1 and 2 are both exponentially distributed, with mean τ1 = λ−1
1 and τ2 = λ−1
2
respectivelya. What are the properties of such a random motion?
The essential quantity is the velocity direction autocorrelation function
C(t0 − t) = hu(t0)· u(t)i = hcos(cid:2)θ(t0) − θ(t)(cid:3)i,
(5)
where u = (cos θ, sin θ) is the direction of motion. Let's introduce the probability densities
pi=1,2(θ, t) to be in mode i with orientation θ at time t, they are governed by
p1(θ, 0) = φ1δ(θ),
p2(θ, 0) = φ2δ(θ),
(6a)
(6b)
where φ1 = λ2/(λ1 + λ2) (resp. φ2 = λ1/(λ1 + λ2)) is the fraction of time spent in mode 1 (resp.
mode 2), and we have taken for initial orientation θ = 0. Now, introducing p(θ, t) = p1(θ, t)+p2(θ, t),
C(t) can be expressed as
∂tp1 = Dr1 ∂2
∂tp2 = Dr2 ∂2
θθp1 + λ2p2 − λ1p1,
θθp2 − λ2p2 + λ1p1,
Z π
−π
C(t) =
dθ p(θ, t) cos θ.
(7)
The system (6) can be solved using Laplace transforms for time and Fourier series for orientation,
yielding
C(s) =
(λ1 + λ2)2 + λ1(s + Dr1) + λ2(s + Dr2)
(λ1 + λ2) [(s + Dr1)(s + Dr2) + λ1(s + Dr2) + λ2(s + Dr1)] ,
where variable s is the Laplace variable. Going back to time domain, this expression translates into
the sum of two exponentials
with the notations
C(t) = c e−γt + c0 e−γ0t,
κ2 = (Dr1 + Dr2 + λ1 + λ2)2 − 4(Dr1Dr2 + Dr1λ2 + Dr2λ1),
κ02 = (Dr1 − Dr2 + λ1 − λ2)2 + 4λ1λ2,
γ = (κ + Dr1 + Dr2 + λ1 + λ2)/2,
γ0 = γ − κ,
c = 1 − c0 = −(Dr1 − Dr2) (λ1 − λ2) + (λ1 + λ2) (λ1 + λ2 − κ0)
.
2κ (λ1 + λ2)
(8)
(9a)
(9b)
(9c)
(9d)
(9e)
aThe value of Dr is thus a Telegraph process.
16
It can be shown that the following inequalities hold
c, c0 (cid:62) 0,
Dr2 (cid:54) γ0 (cid:54) Dr1 < γ,
(10)
showing that both terms are always decaying, and that the slowest relaxation is intermediate between
Dr1 and Dr2.
In the limit Dr1 → ∞, i.e. when all directional persistence is lost in mode 1, all expressions greatly
simplify
γ = Dr1,
γ0 = Dr2 + λ2,
c0 = λ1
λ1 + λ2
,
(11)
where only the first term in the expansion has been retained. In this case, C(t) exhibits a rapid
drop with characteristic time ∼ Dr1
−1, followed by a slowest decay whose constant γ0 = Dr2 + λ2
is independent of Dr1 and whose prefactor c0 is the fraction of time spent in mode 2. On further
assuming that mode 2 involves ballistic motion (Dr2 = 0), then γ0 = λ2, giving access to the mean
duration of mode 2.
In this particular case, the relationship between the biexponential form of
correlation function and the model parameters is simple. This is not so, however, in the general
case, and accordingly we have resorted to a fitting procedure.
Fitting procedure -- experimental data. We use the expressions above (8 & 9) to fit all
experimental correlation functions. Since mode 2 is assumed ballistic, Dr2 = 0. The parameters
Dr1 and τ2, considered as intrinsic properties of the cells, are common to all curves and are thus
heavily constrained. The free parameters remaining for each curve are τ1, and an additional constant
prefactor that allows C(t = 0) to differ from unity.
In practice, we first varied systematically
the values of Dr1 and τ2. For each couple, we fitted the complete set of 15 C(t) curves (5 first
time-windows for each of the 3 density conditions) with the two mentioned free parameters. We
used the value of R2 averaged on these 15 curves to estimate the quality of the fit for a given
−1 = 1.7 min and τ2 = 8.6 min
couple (Dr1, τ2). This yielded unambiguously the optimal values Dr1
(Supplementary Figure 5). Then we measured τ1 by fitting again all the curves using these optimised
fixed parameters.
Supplementary Figure 5: Parameter space for the fitting procedure.
The average R2 computed from 25 experimental curves of correlation functions exhibits a clear peak
−1 = 1.7 min and τ2 = 8.6 min. Here, the time-windows for t (cid:62) 200 min were not considered
at Dr1
in order not to overweight the long-term behaviours. Using the complete set of 39 curves slightly
moves the peak of R2 but then Dr1
−1 = 2 min and τ2 = 10 min remain excellent estimates.
In using the model, we have tacitly assumed that at each time, the population of cells is "equili-
brated" between the two modes. This is reasonable since the residence time in each mode (τ1 and
17
τ2 (min)Dr1−1 (min) 12320151050.9650.970.9750.980.9850.99τ2 = 10 min) remains smaller than the time scale over which λ1 (or the density) significantly varies
(around 100 min).
Supplementary Figure 6: Simulations -- Results of the fitting of the velocity autocorrelation functions
applied to simulations with CEL.
(b) Prefactor c0 of the longest relaxation in the correlation
(a) Proportion of cells in mode 2.
function. Theoretical prediction from φ2 ( -- ) and measurement from the correlation functions (o).
Fitting - simulations with CEL. We computed the proportion of particles in persistent mode
φ2 at all times (direct output from the simulations, Supplementary Figure 6a) and the velocity
autocorrelation functions at various times and for all particle numbers N0. From the latter, we
could extract the parameters from a fit with expression (8). In particular we compared the obtained
c0 values to the theoretical predictions (9e) (Supplementary Figure 6b). They are in close agreement
with each other, showing that the hypothesis of equilibration between the two modes through λ1,
under the control of collisions, is reasonable.
4 Analysis of pair collisions
4.1 Detection of contacts.
Supplementary Figure 7: Radial distribution function, averaged over all the experiments, at various
time points.
18
010020030040000.20.40.60.81t (min)φ2 aN0=20N0=50N0=100N0=200N0=300010020030040000.20.40.60.81t (min)c′b01020304050r(µm)00.511.5g(r)t=0.5ht=1ht=1.5ht=2ht=2.5ht=3ht=3.5ht=4hWe study here the effect of cell-cell contacts on the trajectories. To that end, we need to find a
way to disentangle this effect from other biases related to the occurence of contacts: For instance,
contacts are more likely to happen in dense areas, where the behaviour is known to be different
on average from sparser zones. For that reason, and to avoid increasing the number of spurious
underlying parameters, we only focused on two-body interactions.
A cell-cell contact was defined by two cell positions being closer than a distance dmax. If the cells
remain close longer than one time frame, the collision time t0 is defined as the time at which the
cell-cell distance is minimal. Moreover, since our measurements cannot be instantaneous -- ie they
are based on pieces of trajectories extending over several time frames -- we selected only those
collisions involving two cells that did not encounter any other collision for a time tf ree before and
after t0, hence reducing drastically the number of exploitable data.
Contrary to the simulations, for which the interaction radius is well-defined, the choice of dmax
is not trivial in experiments because cell shapes and sizes are distributed. Thus, to fix dmax we
first measured the radial distribution function g(r) (Supp. Fig. 7). The profiles, computed at
different times, all show a marked peak at r = 10 µm. As a consequence, we chose a slightly larger
dmax = 11 µm to detect reliably the contacts between cells.
4.2 Angular deflection
Supplementary Figure 8: Sketch of the angle definition.
(a) A cell-cell contact is detected in the frame of the picture. (b) The frame is rotated so that the
(O1OO2) axis becomes the new x-axis, where Oi is the position of cell i and O is the barycentre of
the collision. (c) The incidence (resp. scattering) angle θ1/2(−t) (resp. θ1/2(t)) is measured from
the cell's position at −t respective to its position at the time of the contact. (d) The incidence (resp.
scattering) angle separation ∆θi (resp. ∆θf) is the difference between the two directions of motion
of the cell couple.
We studied in details the statistics of incidence and scattering angles, both in simulations and
experiments. To that end, we detected the collisions with tf ree = 3 min before and after the collision.
19
As depicted in Fig. 8 the cell coordinates were rotated so that the (O1OO2) axis is the new x-axis,
where O1 and O2 are the positions of cell 1 and 2 at collision time, and O is the center of [O1O2]. We
then measured the mean angles of motion θi(±t), where i ∈ {1, 2} and t ∈ {1, 2, 3} min, defined as
the angle between (OiXi(t)) and (OOi) and computed ∆θi = θ2(−t)−θ1(−t) and ∆θf = θ2(t)−θ1(t),
the incidence and scattering angles respectively. Trying to decipher the effect of the collision, we
are especially interested in the relative scattering ∆θf as a function of ∆θi. In Fig. 9, we show the
heat maps for the probability densities P(∆θf∆θi).
Supplementary Figure 9: Angular deflection at collision in simulations and experiments.
P(∆θf∆θi) for collisions detected with dmax = 11.2 µm, tf ree = 3 min, at tf /i = ±1 min.
Simulations without (CIL -- ) or with (CIL+) repulsive torque and monomodal (D−1
) or bimodal (CEL+) motion. (b) Experimental data gathered from all experiments.
In simulations without CIL, the probability distribution P(∆θf∆θi) is concentrated around the
main diagonal ∆θf = ∆θi. It means that the relative direction of motion remains unaffected by
the collision, hence that the particles cross "without seeing" each other (in terms of direction of
motion). It is consistent with the simulation rule that the interaction only involves a pushing force
during the contact, but the intrinsic direction of motion is not modified.
(a)
r = 10 min, CEL --
The results are completely different in simulations with CIL (introduced in the form of a repulsive
torque). Most of the probability concentrates in a zone where ∆θi and ∆θf have opposite signs, a
signature of the angular repulsion.
In the experimental data, two distinct behaviors seem to be present: There is both a concentration
of probability around both diagonals, ∆θf = ∆θi denoting crossing, and ∆θf = −∆θi, a signature
of specular reflection. In particular, ∆θf = −∆θi seems more likely for small ∆θi, when the incident
trajectories are close to being parallel, while ∆θf = ∆θi is predominant near k∆θik = π. These
results support the idea that even in the event of a collision, Dictyostelium discoideum cells reorient
smoothly so as to bypass their encounter: at low incidence angle crossing is difficult while almost
specular reflection demands only a slight turn, and at higher incidence it is easier to circumvent
each other. It is different from the usual view of CIL, according to which colliding cells reorient
specifically away along the contact axis.
It could be related to the increased probability for D.
20
discoideum cells to form new pseudopods in the protruded area rather than along the cell body37,
acting together with a CIL-like inhibition of protrusions in the cell-cell contact zone. In any case,
they show that the cell-cell contacts have no aligning effect on the direction of motion, and thus
could not generate a coherent motion that would be responsible for the increase of the colony
spreading rate
4.3 Contact enhancement of locomotion.
Supplementary Figure 10: Increase in cell speed after collisions.
(a) Average speed before and after a single collision, using smoothed trajectories with δt = 1 min and
tf ree = 3 (black), 10 (green), 15 (blue) and 20 min (red). Mean±SEM for n = 6578, 376, 232 and
153 cell pairs respectively. (b) Average normalised speed v/v, where v is the basal speed of a single
cell before and after collisions. Same δt, tf ree and number of contacts as in (a). (c) Logarithm of
the p-value obtained from the Kolmogorov-Smirnov test against the null hypothesis that v(t)/v is
distributed with the same PDF as v(−tf ree)/v. Values above 1 denote when the null hypothesis can
be rejected with more than 95% confidence.
Cell speed. Similarly as for the direction of motion, we measured the instantaneous speed of cells
undergoing a single collision in a frame with free motion tf ree before and after the collision. We
found that on average, the cells exhibit a transient increase of their speed after collisions. Because
of positional noise, this effect is not completely clear using the experimental time frame δt = 20 s.
Yet it is better seen after smoothing the trajectories over δt = 1 min, and varying tf ree from 3 min to
20 min: tf ree = 20 min provides a view of the long-term dynamics; with shorter tf ree, the complete
time-frame is not accessible, but the trend for hvi(t) is confirmed with much more statistics (up to
6578 cell-cell contacts measured with tf ree = 3 min).
The cells accelerate for 7− 8 min following collisions, reaching an average speed approximately 20%
higher than the "basal" average speed, and then the speed decreases back to its basal value at
a similar rate. When the speed of each single cell is normalised by its own basal speed v (i.e.
its mean speed prior to the collision), one even gets a 30% increase on average. Although these
changes are relatively small compared to the standard deviation of the speed distribution in the cell
population, a statistical analysis using the Kolmogorov-Smirnov test shows that the distribution
differs significantly from the distribution at t = −tf ree only for ∼ 12 min after collisions.
Persistence time. Using again smoothed trajectories with δt = 1 min to make the patterns
of evolution more apparent, we repeat the analysis on the CME computed for ∆t = 5 min (see
Methods). This revealed a CME drop around the contacts, as well as a subsequent transient increase
(Supp. Fig. 11a-b). Both effects seemingly arise from significant changes in the distribution of the
21
-20020t−tcol(min)012345hvi(µm.min−1)atfree=3mintfree=10mintfree=15mintfree=20min-20020t−tcol(min)0.811.21.4hv/vib-20020t−tcol(min)0246log(p)/log(0.05)cSupplementary Figure 11: Increase in CME after collisions.
(a) Average CME before and after a single collision, using smoothed trajectories with δt = 1 min
and tf ree = 10 (green), 15 (blue) and 20 min (red). Mean±SEM for n = 376, 232 and 153 cell
pairs respectively. (b) Average normalised CME CME/CME, where v is the basal speed of a single
cell, before and after collisions. Same δt, tf ree and number of contacts as in (a). (c) Logarithm of
the p-value obtained from the Kolmogorov-Smirnov test against the null hypothesis that CME(t) is
distributed with the same PDF as CME(−tf ree). Values above 1 denote when the null hypothesis
can be rejected with more than 95% confidence.
CME (Supp. Fig. 11c). The depression is probably related to changes in direction during the
contact. The increase, between 10 and 15% for single cells on average, shows that the persistence
of the motion is enhanced for at least a few minutes after a cell-cell contact.
Supplementary Figure 12: Single-cell-normalised CME: comparison of various simulation conditions
with experimental data. Mean±SEM, tf ree = 15 min.
Sensitivity of CME measurements to the presence of CEL. While in the case of the speed
the measurement is direct, the CME is only a proxy to estimate whether the trajectories are more
or less persistent on a given time-scale ∆t. As a consequence, we also tested our CME-based
analysis of collisions on the simulation data. It makes apparent that a drop in CME around 10% is
the signature of CIL (in the sense of turning upon contact), while a transient increase in CME of
10 -- 20% is observed only in the presence of CEL (Supp. Fig. 12).
Finally, the experimental curve aligns closely with that of the CIL+/CEL+ simulations (Supp.
Fig. 12). Without proving that the cells follow the precise rules implemented in the simulations, it
22
-20020t−tcol(min)0.50.60.70.8CMEtfree=10mintfree=15mintfree=20min-20020t−tcol(min)0.911.11.2CME/CME-20020t−tcol(min)02468log(p)/log(0.05)c-10-50510t-tcol (min)0.850.90.9511.051.11.151.2CME/CMECIL-/CEL-CIL+/CEL-CIL+/CEL+CIL-/CEL+Expeshows that CIL and CEL seem to be necessary to account for the experimental collision data.
Supplementary movies
Supplementary movie 1 Long-time spreading of a colony with high cell density.
Supplementary movie 2 Short-time (200 min) dynamics of a colony with low cell density. N0 =
18.
Supplementary movie 3 Short-time (200 min) dynamics of a colony with high cell density.
N0 ' 275.
Supplementary movie 4 Simulation of active particles at low density undergoing CEL. N0 = 20.
Supplementary movie 5 Simulation of active particles at high density undergoing CEL. N0 =
200.
23
|
1111.2707 | 1 | 1111 | 2011-11-11T11:18:46 | Measuring the viscous and elastic properties of single cells using video particle tracking microrheology | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.CB"
] | We present a simple and \emph{non-invasive} experimental procedure to measure the linear viscoelastic properties of cells by passive video particle tracking microrheology. In order to do this, a generalised Langevin equation is adopted to relate the time-dependent thermal fluctuations of a bead, chemically bound to the cell's \emph{exterior}, to the frequency-dependent viscoelastic moduli of the cell. It is shown that these moduli are related to the cell's cytoskeletal structure, which in this work is changed by varying the solution osmolarity from iso- to hypo-osmotic conditions. At high frequencies, the viscoelastic moduli frequency dependence changes from $\propto \omega^{3/4}$ found in iso-osmotic solutions to $\propto \omega^{1/2}$ in hypo--osmotic solutions; the first situation is typical of bending modes in isotropic \textit{in vitro} reconstituted F--actin networks, and the second could indicate that the restructured cytoskeleton behaves as a gel with "\textit{dangling branches}". The insights gained from this form of rheological analysis could prove to be a valuable addition to studies that address cellular physiology and pathology. | physics.bio-ph | physics | Measuring the viscous and elastic properties of single cells using
video particle tracking microrheology.
Rebecca L. Warren, Manlio Tassieri∗, Xiang Li, Andrew Glidle, Allan Carlsson, Jonathan M. Cooper
Division of Biomedical Engineering, School of Engineering, University of Glasgow,
1
Glasgow, U.K.
∗ E-mail: [email protected]
Abstract
We present a simple and non-invasive experimental procedure to measure the linear viscoelastic properties
of cells by passive video particle tracking microrheology.
In order to do this, a generalised Langevin
equation is adopted to relate the time -- dependent thermal fluctuations of a bead, chemically bound to the
cell's exterior, to the frequency -- dependent viscoelastic moduli of the cell. It is shown that these moduli
are related to the cell's cytoskeletal structure, which in this work is changed by varying the solution
osmolarity from iso -- to hypo -- osmotic conditions. At high frequencies, the viscoelastic moduli frequency
dependence changes from ∝ ω3/4 found in iso -- osmotic solutions to ∝ ω1/2 in hypo -- osmotic solutions; the
first situation is typical of bending modes in isotropic in vitro reconstituted F -- actin networks, and the
second could indicate that the restructured cytoskeleton behaves as a gel with "dangling branches". The
insights gained from this form of rheological analysis could prove to be a valuable addition to studies
that address cellular physiology and pathology.
Introduction
As many authors have noted, the mechanical properties of a cell's cytoskeleton can influence factors
such as growth, apoptosis, motility, signal transduction and gene expression [1]. Related to this, there
is a desire to be able to provide a rheological interpretation of the cell's viscoelastic response that has
the potential to yield quantitative information on the cell's cytoskeletal structure. Consequently, in this
work, we have developed a means of using passive video particle tracking microrheology measurements
to quantitatively measure changes in the viscoelastic properties of a cell as a consequence (in this case)
of simple changes in its external environment, i.e. subjecting a cell to a hypo -- osmotic shock.
1
1
0
2
v
o
N
1
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
0
7
2
.
1
1
1
1
:
v
i
X
r
a
2
Osmotic regulation and the transport of osmotically active molecules are fundamental to both metabolic
processes and homeostasis of cells and require precise regulation and maintenance of intracellular water
[2]. The ability of many cells to regulate their volume, via internal restructuring, in response to osmotic
changes in their environment is an essential component of normal cellular function. This regulatory vol-
ume change is linked to a reorganisation of the cytoskeletal actin networks [3, 4]. Exposure to a hypotonic
solution typically induces an initial rapid swelling followed by a shrinkage of the cell that occurs over
a slower timescale of several minutes. The modulation of the actin dynamics and polymerisation that
accompanies these regulatory volume changes has been shown in various cell types by methods such as
DNase I inhibition assay and fluorescence measurements of phalloidin -- labelled actin [3].
In this work we have developed a video particle -- tracking tool to study the microrheology of cells,
using a Jurkat lymphocyte cell line as a model system. In general, lymphocytes have been shown to have
a response to hypo -- osmotic solutions that is influenced by tyrosine kinase activity [5] and cytoskeleton
(F -- actin) participation in ion channel activation [6]. Furthermore, the interest in T -- cells stems from
their important role in the regulation of immune responses. In particular, Jurkat cells, as used in this
study, are CD4+ T -- lymphoma cells that are often utilised as models to understand T -- cell signalling [7]
and HIV-1 dissemination in viral pathogenesis [8].
The linear viscoelastic properties of a material can be represented by the frequency -- dependent dy-
namic complex modulus G∗(ω), which provides information on both the viscous and the elastic nature
of the material (i.e. on how the matter dissipates and stores the energy transferred to it) at different
frequencies ω; it is defined as the ratio between the Fourier transforms (denoted by the symbol " ") of
the applied stress σ(t) (which is proportional to the applied force) and the resulting strain γ(t) (which is
proportional to the material/cell deformation):
G∗(ω) =
≡
σ(t)
γ(t)
(cid:82) +∞
(cid:82) +∞
−∞ σ(t)e−iωtdt
−∞ γ(t)e−iωtdt
(1)
where i is the imaginary unit (i.e. i2 = −1).
The standard method of measuring G∗(ω) is based on the imposition of an oscillatory stress σ(ω, t)
given by a formula such as:
σ(ω, t) = σ0 sin(ωt)
(2)
where σ0 is the amplitude of the stress function, ω is the frequency and t the time. The resulting
oscillatory strain γ(ω, t) then has the formula:
γ(ω, t) = γ0 sin(ωt + δ(ω))
3
(3)
where γ0 is the strain amplitude and δ(ω) is the frequency -- dependent phase shift between the stress and
the strain. The amplitudes of the complex modulus in -- phase and out -- of -- phase components are propor-
tional to the ratio between amplitudes of the stress and of the strain, with constants of proportionality
defining the storage (elastic) G(cid:48)(ω) and the loss (viscous) G(cid:48)(cid:48)(ω) moduli, respectively:
G∗(ω) = G(cid:48)(ω) + iG(cid:48)(cid:48)(ω) =
σ0
γ0
cos(δ(ω)) + i
σ0
γ0
sin(δ(ω))
(4)
For example, in the case of a purely elastic solid, the stress and the strain are in phase (i.e. Hooke's law,
the material deformation is directly proportional to the applied force) and δ(ω) = 0; whereas, for a purely
viscous fluid, such as water or glycerol, δ(ω) = π/2. For complex solids (e.g. gels, cells) or viscoelastic
fluids (e.g. blood, saliva) δ(ω) would take any value between these limits (i.e. 0 (cid:54) δ(ω) (cid:54) π/2) depending
on the frequency at which the force (stress) is applied.
The aim of this article is to present a straightforward and non -- invasive procedure for measuring the
in vivo linear viscoelastic properties of single cells via passive video particle tracking microrheology of
single beads attached to the cells' exterior. This method has advantages over both complicated active
microrheology techniques, where complex experimental set -- ups are necessary to exert an external force
for performing stress -- controlled measurements; and invasive passive video particle tracking of submicron --
probes embedded (either via endocytosis or micropipette injection) within the cell's cytoskeleton (e.g.
[9, 10, 11, 12, 13, 14, 15]. In particular, the procedure consists of measuring the thermal fluctuations of a
bead chemically bound to the cell's exterior (Figure 1), for a sufficiently long time. A generalised Langevin
equation was adopted to relate the time -- dependent bead trajectory, (cid:126)r(t), to the frequency -- dependent
moduli of the cell. Notably, the procedure presented here represents an alternative methodology that
can be extended to many experimental formats and provides a simple addition to existing cellular phys-
iology studies (e.g. those monitoring cell pharmacological response). Indeed, when compared to single
cell viscoelasticity assays such as magnetic tweezers, atomic force microscopy and optical stretcher, as
employed in [9, 10, 11, 12, 13, 14, 15], our method has the advantage of revealing the changes of the cell's
viscoelastic properties over a wide range of frequencies (here from ∼0.6 Hz up to ∼600 Hz), to a high
level of accuracy, whilst it experiences an induced physiological process.
4
Figure 1. (Left) Top-view of a 5 µm diameter silica bead bound to the surface of a Jurkat
cell. (Centre) Schematic side-view of the left image. (Right) The thermal fluctuations of a 5 µm silica
bead chemically bound to the surface of a Jurkat cell. Images were analysed in real -- time and the
coordinates of the bead's centre of mass were stored directly on the random -- access memory (RAM) of
the computer, at time intervals down to 1/600s, using our own suite of image analysis software written
in LabVIEW [23].
Analytical model
The equation describing the pseudo Brownian fluctuation of the randomly varying bead position (cid:126)r(t)∀t
can be derived by means of a generalised Langevin equation, which was originally derived to describe
the Brownian motion of a freely diffusing particle suspended in a purely viscous fluid [16] and represents
Newton's second law as written for stochastic processes:
(cid:90) t
m(cid:126)a(t) = (cid:126)fR(t) −
[ζc(t − τ ) + ζs(t − τ )](cid:126)v(τ )dτ
(5)
0
where m is the mass of the particle, (cid:126)a(t) is its acceleration, (cid:126)v(t) is the bead velocity and (cid:126)fR(t) is the usual
Gaussian white noise term that models stochastic thermal forces acting on the particle. The integral
term represents the total damping force acting on the bead, which, based on the superposition principle,
incorporates two generalised time -- dependent memory functions ζc(t) and ζs(t) that are representative
of the viscoelastic nature of the cell and the solvent, respectively. These memory functions are directly
related to the materials' complex modulus as shown hereafter.
In the case of the solvent, which in this work is the media surrounding the cell, the relationship
between memory function and complex modulus is straightforward. Indeed, to a first approximation,
using the assumptions adopted by Mason and Weitz [17] when studying the motion of thermally excited
0246810-100-50050100 r(t) [nm]Time [s]5
free particles, at thermal equilibrium the complex viscoelastic modulus of the solvent is related to the
memory function ζs(t) by the expression: G∗
is the Fourier transform of ζs(t). In the case of the cell, we assume a similar relationship between G∗
s(ω) = iω ζs(ω)/6πR, where R is the bead radius and ζs(ω)
c (ω)
and ζc(t) to that given above for the solvent, but with a different constant of proportionality that we will
call β; i.e., G∗
c (ω) = iω ζc(ω)/β. Note that, β may vary for different cells because it depends on (i) the
cell radius (Rc), (ii) the number of the chemical bonds between the bead and the cell, (iii) the contact
area between the cell and the glass coverslip, and (iv) the relative position of the bead with respect to
both the cell's equatorial plane and the glass coverslip. In this study we will only focus on the changes
in cell dynamics due to variations in osmolarity. This can be achieved since all the above experimental
parameters, with the exception of Rc, are (in good approximation) time -- independent, and will not change
significantly during the course of a set of measurements, as shown in Figure 4 for measurements performed
on cells in iso -- osmotic condition. In addition, it is important to highlight that the cells maintained a
spherical shape well beyond the duration of the experiment. Moreover, the cells adhered to the coverslip
and no drift was observed during the experiment. Finally, given that we measure an increase of the cell
radius (Rc) of (cid:54) 5% for the hypo -- osmolarity experiments described below, we also make the assumption
that this small change does not affect the dynamics of the system appreciably.
We now describe how the thermal fluctuations of a bead, chemically bound to a cell, can be investigated
to determine the viscoelastic properties of the cell through the analysis of the time dependence of the
bead's mean -- square displacement (MSD)(cid:10)∆r2(τ )(cid:11) ≡(cid:68)
[(cid:126)r(t + τ ) − (cid:126)r(t)]2(cid:69)
, in the radial direction of the
t
cell; where t is the absolute time and τ is the lag -- time (i.e. time interval). The average is taken over
all initial times t. Under these circumstances, it is an easy step to show that, at thermal equilibrium,
Equation 5 can be reorganised to express the viscoelastic moduli of the cell as function of the Fourier
transform of the mean -- square displacement:
G∗
c (ω)
G(cid:48)
0
=
1
iω Π(ω)
+
mω2
βG(cid:48)
0
− 6πRG∗
βG(cid:48)
0
s(ω)
(6)
where G(cid:48)
(i.e. cell plus solvent), which in this work is G(cid:48)
0 is the limiting value, for vanishingly low frequencies, of the elastic modulus of the whole system
c (ω) for ω → 0; and Π(ω) is the Fourier transform
of the normalised mean -- square displacement Π(τ ) =(cid:10)∆r2(τ )(cid:11) /(cid:10)r2(cid:11), introduced by Tassieri et al. [18]
0 ≡ G∗
in the study of the thermal fluctuations of an optically trapped bead suspended in a viscoelastic fluid.
The term(cid:10)r2(cid:11) is the time -- independent variance of (cid:126)r(t) measured for a sufficiently long time. Moreover,
6
it has been shown [18] that for a constrained bead (e.g. one that is optically trapped or, as in this work,
chemically bonded to the cell) Π(τ ) = 1 at large time intervals. It is important to highlight that (as in
the case of an optically trapped bead) the term βG(cid:48)
0 (representing the coefficient of the elastic restoring
force) can be determined by means of the principle of equipartition of energy, which in one dimension is
written as:
kBT = βG(cid:48)
0(cid:104)r2(cid:105)
(7)
Usefully, two further simplifications can be made to Equation 6 because (a) for micron-sized silica
beads, the inertia term mω2 is negligible up to frequencies on the order of M Hz and (b) for solvents having
frequency -- independent viscosity η (e.g. water), G∗
viscoelastic properties of the cell (scaled by G(cid:48)
s(ω) simplifies to iωηs. Thus, Equation 6 provides the
0) over a range of frequencies that is limited at the top end
by the rate of acquisition of images to determine the bead position, and at the bottom end by a cutoff
frequency given by ωcutof f = βG(cid:48)
0/η0, where η0 is the limiting value, for vanishing frequencies, of the
viscosity of the whole system.
Finally, in order to evaluate the Fourier transform in Equation 6 we adopt the analytical method
introduced by Evans et al. [19] (i.e. Eq. (10) of Ref. [19]), which is applied directly to the experimental
data points and has the advantage of removing the need for Laplace and inverse -- Laplace transformations
of experimental data [20].
Experimental details
Anti-CD4+ (Invitrogen) was attached to 5µm carboxylate functionalised silica beads (Bangs Laboratories
Inc) using a method based on a previously described protocol [21]. The adaptation of this method involved
first binding a short polyethylene glycol chain to the beads so as to provide a flexible linker on to which
the anti-CD4+ could be bound.
In detail, the microspheres were washed (3x) with 0.1 M N a2CO3 buffer (pH 9.6) using centrifugation
between washes to remove the supernatant liquid. This ensured that all the carboxylate groups were
deprotonated. Following the final wash, the beads were resuspended in RO water and centrifuged again.
After removal of the water, the remaining pellet was resuspended in freshly prepared 0.02 M sodium phos-
phate buffer (pH 6) containing 2 mM 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide (EDC, Sigma) and
7
5 mM sulfo-N-hydroxysuccinimide (sulfo-NHS, Thermo-Fisher). The microsphere solution was then ag-
itated for 15 minutes to allow the sulfo-NHS to activate the carboxylate groups on the surface of the
microspheres. The solution was centrifuged, then the supernatant removed and replaced by a 10 mM solu-
tion of the flexible linker, O-(2-Aminoethyl)-O'-(2-carboxyethyl)polyethylene glycol 3,000 hydrochloride
(Aldrich) in pH 7.4 phosphate buffer. The activated microspheres were then agitated in the flexible linker
solution for 2 h at 4oC, before further centrifugation and washing with pH 7.4 buffer. The anti-CD4+
was then bound to the carboxylate terminals of the PEG linker from a 1 mg/ml solution of anti-CD4+,
using the same EDC/sulfo-NHS activation procedure outlined above.
The Jurkat cells were obtained from ATCC (clone E6-1, TIB-152) and maintained in RPMI 1640
(1x) media containing L-Glutamine and 25mM HEPES [Gibco, Invitrogen]. Prior to the experiment they
were suspended in PBS [Gibco, Invitrogen] with an osmolarity of 275-304 mOsm, then mixed with the
functionalised beads and allowed to equilibrate for ∼10min.
Beads were optically trapped and moved to the equatorial plane of the cell, where they were held for
a sufficient time (∼2 min) until the binding process was completed. Trapping was achieved by using a
CW Ti:sapphire laser system (M Squared, SolsTiS) which provides up to 1 W at 830 nm. Holographic
optical traps were created via the use of a spatial light modulator (Boulder XY series) in the Fourier
plane of the optical traps [22]. The microscope was equipped with a 100x 1.3NA (Zeiss, Plan-Neofluor)
objective lens, which was used both to focus the trapping laser beam and to visualise the particles via a
Prosilica PCI-6713 camera. Samples were mounted on a motorised microscope stage (Prior Pro-Scan II).
The beads' trajectories were measured and stored on a personal computer in real-time, at frequencies up
to ∼kHz (depending on the size of the analysed region of interest (ROI) within the camera field of view),
using our own suite of image analysis software written in LabVIEW [23].
Results and Discussion
In order to validate the analytical model introduced in this work, we measured the dynamic response of
an analytical system that could be considered as an ideal viscoelastic medium with known storage and
loss moduli; i.e. an optically trapped bead suspended in a Newtonian fluid. Indeed, for such a system,
the elastic modulus must be proportional to the optical tweezers' (OT) trap stiffness (κ), this being the
only elastic component present in the system (i.e. the ensemble of Newtonian fluid, bead and OT); and
the viscous modulus must be linearly proportional to the frequency (ω) with constant of proportionality
given by the viscosity of the Newtonian fluid (η).
8
Figure 2. (A) The mean-square displacements vs. lag-time of optically trapped beads in
Newtonian fluids. The square and the triangle symbols refer to measurements performed with 2 µm
diameter silica bead in Glycerol/water mixtures at 10% and 20% w/w, having viscosity values of 1.15
mP as and 1.54 mP as, and trap constants of 0.85 µN/m and 0.81 µN/m, respectively; The circle
symbols refer to a measurement performed with 5 µm diameter silica bead in water having viscosity
value of 0.89 mP as and trap constants of 1.68 µN/m (note that for this measurement only we used an
off-line faster camera with internal memory; not suitable for continuous acquisition). The inset shows
the same data as above but with the abscissa scaled by κ/(6πηR). (B) The normalised viscoelastic
moduli vs. the normalised frequency for all three analytical systems described above.
In Figure 2B we present the normalised viscoelastic moduli versus the normalised frequency for three
systems made up with diverse combinations of Newtonian fluids and beads having different viscosities
and size, respectively. For each system, the moduli have been derived by the analysis of the thermal
c (ω) in Equation 6 represents the viscoelastic nature of the analytical system;
fluctuations of an optically trapped bead by means of Equation 6 via the relative MSD shown in Figure 2A.
In this case, the term G∗
the term G∗
c (ω)] (and therefore it does not need to be
0 ≡ κ is determined by means of the principle of equipartition of
accounted for in Eq. 6); the term βG(cid:48)
energy with β ≡ 6πR. In summary, for both moduli, the normalising factor is given by κ/(6πR); whereas,
s(ω) simplifies to iωη and is coincident with (cid:61)[G∗
for the abscissa, the scaling factor is given by a cut-off frequency defined as κ/(6πRη), which also works
in the time-domain as shown in the inset of Figure 2A. In Figure 2B it is interesting to highlight that
at low frequencies (i.e. up to the cut-off frequency) the elastic modulus G(cid:48)(ω) is actually normalised
to 1, as expected; whereas, at high frequencies (i.e. for ω > κ/(6πR)) G(cid:48)(ω) vanishes towards zero. The
latter phenomenon is itself an interesting feature that is currently an object of discussion in literature
(A) (B) (A) (B) 9
[24, 25, 26]. In particular, it is argued that the analytical models adopted so far for data analysis of
the thermal fluctuation of an optically trapped bead suspended into a Newtonian fluid are unable to
provide the high frequency elastic component of the optical tweezers (OT), because they would "suffer
from systematic truncation errors at high frequencies". In contrast, we believe that the absence of the
optical tweezers' high frequency elastic component is an intrinsic feature of the system made up by the
ensemble of (Newtonian) fluid, bead and OT [27]; on which we will provide a detailed explanation, but
in a different work. On the other hand, the viscous modulus G(cid:48)(cid:48)(ω) grows linearly with the frequency
(thus in a Log-Log plot it is represented by a line with unity slope) over the entire experimental frequency
window, with constant of proportionality given by the viscosity of the Newtonian fluid, as expected. Note
that the results shown in Figure 2 are in good agreement with those published in literature [24, 25, 26]
for an optically trapped bead in water, but they used a different analytical method for data analysis and
they overlooked the scaling process introduced in this work.
Having demonstrated the validity of our analytical model, we now investigate its applicability to
cellular physiology studies by focussing on changes in the cells' microrheology due to changes in osmolarity
of the surrounding solution.
It is anticipated that there are two major physiological processes that accompany the changes in
osmolarity of the solution surrounding the Jurkat cell: (I) an actin cytoskeleton reorganisation and (II)
an increase of the cell volume as the solution becomes hypo -- osmotic. In order to quantify the changes
in cell volume, we assumed that the cells were spherical, Figure 1. In particular, Figure 3 shows that
the relative radius changes are < 6% when the surrounding PBS solution was made hypo-osmotic by
addition of 10% v/v distilled water. The data of Figure 3 also shows that, after an initial swelling, the
cells contract to a still swollen state that is ∼5% larger than when they were in isotonic PBS. Based on
these observations, which are in good agreement with those already existing in literature (e.g. [28]), we
assume that these small radius changes do not affect the dynamics of the system appreciably.
To study how the in vivo viscoelastic properties of Jurkat cells vary with the osmolarity of the solution,
we have adopted the model described in the Analytical Section. In particular, analysis of different time
regimes in the time dependent normalised mean-squared displacement, Π(τ ), has the potential to reveal
information both on the pure elastic component of the cell (at long time intervals or low frequencies) and
on the fast dynamics occurring at small length scales (e.g. those related to the transverse bending modes
of single actin filaments in the cytoskeleton).
10
Figure 3. Averaged relative cell radius changes vs. time as a consequence of the change in
osmolarity of the cell solution; from iso-osmotic conditions in PBS buffer towards a
hypo-osmotic condition obtained by adding 10% v/v of distilled water to the initial
solution. The osmotic shock causes a rapid swelling of the cells, followed by a partial recovery of the
cells' volume. These results are in good agreement with those reported in literature, e.g. [28].
Elastic properties of the cell derived from Π(τ )
The first result that can be obtained through the analysis of the thermal fluctuations of the bound bead
(Figure 1) is the relative change of the cell's low frequency elastic plateau modulus G(cid:48)
c(0), Figure 4.
This can be evaluated from the time -- independent variance (cid:104)r2(cid:105) of the constrained bead in the radial
direction of the cell, by adopting the principle of equipartition of energy as in Equation 7. In particular,
although the absolute value of G(cid:48)
change between different values of βG(cid:48)
c(0) can not be determined because of the unknown factor β, the relative
c(0), obtained from n sequential measurements performed on the
same cell, differ from each other only because of both the small increase of the cell radius (measured as
∼5%) and the variation of G(cid:48)
c(0). Thus, from Figure 4 it is clear that a change of 10% in osmolarity
towards hypotonicity, which occurs at t ∼9min as a result of adding distilled water to the iso-osmotic
solution, induces a temporary increase of βG(cid:48)
this change in the low frequency elastic modulus dependent term, βG(cid:48)
c(0) that lasts for approximately 15-30min. At its greatest,
c(0), is over 3 fold in magnitude.
The explanation for such a large increase may in part be due to the change in radius; however, if this
were solely responsible, it would be expected that βG(cid:48)
c(0) would remain significantly above its original
0510152011.011.021.031.041.051.06Time [min]Rc/Rc011
Figure 4. The absolute value of βG(cid:48)
symbols) and when the solution is made 10% hypotonic (square symbols) The arrow indicates
the time at which the isotonic solution is made hypotonic. The dotted line is a guide for the eye, so
that the later points in the graph can be compared to the first points measured in isotonic condition.
c(0) measured in isotonic (PBS) condition (circle
value, as the cells do not regain their original radius. Thus, since βG(cid:48)
c(0) does return to close to the
original value (at t ∼25 min), consideration should be given to the temporary increase in electrostatic
intermolecular interactions within the cell due to the reduction in ionic strength of the solution (i.e.
a reduction in screening of the charges on cytoskeleton molecules by solution based ions). This latter
phenomenon would also alter the cell's viscoelastic properties and any resulting actin cytoskeleton re-
arrangement may be seen as part of an attempt of the cell to re -- equilibrate the solution osmolarity, as
discussed below.
Evidence for an increase in intermolecular interactions in the cytoskeleton being the cause of the
observed increase in βG(cid:48)
modulus (G(cid:48)(0)) of an ideal cross-linked polymer network can be written:
c(0) comes from studies of polymer gels [29, 30, 31], for which the elastic plateau
G(cid:48)(0) ∼= ψkBT
(f − 2)
f
(8)
where kB is the Boltzmann constant, T is the absolute temperature, and ψ and f are the number
of network strands and cross -- links per unit volume, respectively (see Figure 5). Here, an increase in
12
number of strands and/or number of cross -- linking interactions causes an overall increase in G(cid:48)(0). Note
that although it is appreciated that there may be more order associated with macromolecules within the
cell's cytoskeleton, the generalised polymer network of Figure 5 nevertheless has many similarities with
the complex structure of the cytoskeleton. For example, it is well known that all cells possess a host of
actin -- binding proteins that can align actin filaments into bundles and then cross-link filaments or bundles
into networks [32]. Bundled actin and scruin in vitro networks act similarly to cross -- linked networks and
show a strong increase in elasticity as the degree of bundling increases [33].
Figure 5. Schematic view of a randomly cross-linked network with "dangling branches"
defects. Circles represent cross-linking junctions and arrows denote attachments to the
macroscopic network. The "dangling branches" do not contribute to the gel elasticity because they
cannot bear stresses; in contrast, they can dissipate energy by friction with the solvent.
Thus, by analogy with the models for a polymer network, a >3 fold increase in the cell elastic plateau
modulus could be explained by an increase of intermolecular interactions (i.e. electrostatic interactions)
that would have the effect of forming temporary cross -- linkages and reshaping the isotropic cytoskeleton
structure. This scenario is in good agreement with confocal microscopy results already reported for
multiple cell lines and the observations that actin forms bundles in aortic bovine endothelial cells after
hypotonic exposure [34].
Cross-linksDangling ends13
Fast dynamics of the cell derived from Π(τ )
In order to study the high frequency behaviour of cells, we have used Equation 6, which directly relates
the normalised mean-square displacement, Π(τ ), of the bead to the cell complex modulus, G∗
c (ω). In
particular, we have applied Equation 6 to the Π(τ ) derived from measurements that were collected at
fixed time intervals after the solution was made hypotonic. In Figure 6, these are compared with the
normalised MSD of the same cell in phosphate-buffered saline (PBS) solution, prior to hypotonic exposure.
Note that in order to more easily discern visually the changes in Π(τ ), caused by the hypotonicity, the
normalised MSD for the cell in PBS is plotted in each of the panels.
The elastic (G(cid:48)
c(ω)) and viscous (G(cid:48)(cid:48)
c (ω)) components of the complex modulus, G∗
c (ω), derived from
the measurements shown in Figure 6 are shown in Figure 7 (scaled by the plateau modulus G(cid:48)
c(0)). It
is clear that, in PBS (Figure 7A), the high frequency behaviour of the cell's elastic and viscous moduli
both show a frequency dependence of ∝ ω3/4, which is characteristic of isotropic in vitro reconstituted
actin filament solutions [35, 36].
As the solution is made hypo -- osmotic, the cell cytoskeleton starts to reorganise and the frequency
behaviour of the moduli drastically change as (sequentially) shown in Figure ??.
In particular, after
∼10min in the hypo -- osmotic solution both moduli tend to assume a high frequency behaviour ∝ ω1/2
(see Figures 7C and 7D). This change in the degree of elasticity may be explained by two alternative
(and possibly concomitant) processes: (i) an increase in cytoskeletal tension as response to a stretching
force caused by the cell swelling and (ii) the consequent formation of "dangling branches" [29, 30, 31]
(Figure 5) generated during the reorganisation process of the cell cytoskeleton (including those possibly
obtained from the rupture of the cytoskeletal protein filaments because of stress overload). However,
whilst the increase in cytoskeletal tension would explain both the initial increase of the cell stiffness
and the change in the frequency scaling laws of the moduli (i.e. from ∝ ω3/4 to ∝ ω1/2), it would fail
to explain the softening process occurring during the cell volume re -- equilibration, which ends with a
breakdown, at high frequency, of the cell's elastic modulus (Figures 7G). This latter phenomenon could
be simply justified by an excess of "dangling branches". Indeed, from a rheological point of view, these
would only contribute to the capability of the cell to dissipate energy but not to the cell elasticity, since
they cannot bear stress and hence do not contribute to the cell's rigidity.
Finally, it is important to highlight that the results that have been obtained here using a video particle
tracking method are in good agreement with those presented in the review written by Papakonstanti and
Stournaras [3], where a set of assessments of actin cytoskeleton dynamics and actin architecture in cell
volume regulation are summarised. However, none of the techniques reviewed in that work [3] are able
to provide quantitative information on the cell viscoelasticity, as is the case here.
14
Figure 6. The normalised mean -- square displacement vs. lag -- time of a 5 µm diameter silica
bead chemically bound to a Jurkat cell in iso -- osmotic (PBS) solution (square symbols)
and in hypo-osmotic solution (circle symbols) after the addition of 10% v/v distilled water
to the PBS buffer and measured at time intervals (∆t) of (A) 2 min, (B) 7 min, (C) 12
min, (D) 18 min, (E) 23 min, (F) 28 min, respectively. At short time intervals, the Π(τ ) has the
potential to reveal the dynamics occurring at molecular level (as shown in Figure ??); whereas, at large
lag -- times it provides information on the stiffness of the whole cell (as shown in Figure 4).
15
c(ω)) and the imaginary (loss, G(cid:48)(cid:48)
0 vs. frequency of a Jurkat cell in iso-osmotic (PBS) solution
Figure 7. The real (storage, G(cid:48)
modulus (G∗
c (ω)) scaled by G(cid:48)
(A) and in hypo-osmotic solution (B-G) after the addition of 10% v/v distilled water to
the PBS buffer and measured at time intervals (∆t) of (B) 2 min, (C) 7 min, (D) 12 min,
(E) 18 min, (F) 23 min, (G) 28 min, respectively. The moduli have been evaluated by using
Equation 6 on the normalised mean-square displacement data shown in Figure 6. The lines are guides
for the gradients. Note that, the scatter in the data at high frequencies and the ripples in the low
frequency portions of the curves in Figure 7 are due to the analytical method that we have adopted [19]
for performing the Fourier transforms of the normalised MSDs. As explained in detail in Ref. [37], this
c (ω)) parts of the complex
method works directly on the experimental data points (i.e. (cid:8)τk, Πk
(cid:9), where k = 1..N ) and therefore
preserves genuine experimental noise.
16
Conclusions
In summary, we have presented a straightforward and non -- invasive experimental procedure, coupled
with a new analytical method to interpret the data, which leads to quantitative determination of the in
vivo viscoelastic properties of cells in the frequency domain. The method has the potential to monitor
the internal dynamics and re -- organisation of the actin cytoskeleton up to frequencies on the order of
kHz, representing a valuable addition to studies that address cellular physiology and pharmacological
response. Indeed, in this work we report, for the first time, the high frequencies (up to ∼600Hz) changes
of the Jurkat cells' viscoelastic spectrum from ∝ ω3/4 to ∝ ω1/2, as response to a change in osmolarity
of the solution. The rheological interpretation of the results gives a direct insight of the cell cytoskeleton
structure and its re -- organisation.
In the future, it is envisaged that these interpretations could be
coupled with advanced molecular biology techniques to resolve the detailed interactions underlying these
rheological changes and that faster dynamics could be studied by means of a quadrant photo -- diode based
tracking system
Acknowledgments
We thank Mike Evans for helpful conversations. M.T. is Research Fellow of the Royal Academy of
Engineering/EPSRC; research program title: "Rheology at the Microscale: New Tools for Bio-analysis".
We are grateful to EPSRC and BBSRC for supporting this work through grants EP/F040857/1 and
BB/C511572/1, respectively, and to the DTC in Proteomic and Cell Technologies (EPSRC) for funding
RLW.
References
[1] Chicurel ME, Chen CS, Ingber DE (1998) Cellular control lies in the balance of forces. Curr Opin
Cell Biol 10: 232-9.
[2] Nagy T, Balasa A, Frank D, Rab A, Rideg O, et al. (2010) O-glcnac modification of proteins affects
volume regulation in jurkat cells. Eur Biophys J 39: 1207-17.
17
[3] Papakonstanti EA, Stournaras C (2007) Actin cytoskeleton architecture and signaling in osmosens-
ing. Methods Enzymol 428: 227-40.
[4] Ebner HL, Cordas A, Pafundo DE, Schwarzbaum PJ, Pelster B, et al. (2005) Importance of cy-
toskeletal elements in volume regulatory responses of trout hepatocytes. Am J Physiol Regul Integr
Comp Physiol 289: R877-90.
[5] Lepple-Wienhues A, Szab`o I, Laun T, Kaba NK, Gulbins E, et al. (1998) The tyrosine kinase p56lck
mediates activation of swelling-induced chloride channels in lymphocytes. J Cell Biol 141: 281-6.
[6] Levitan I, Almonte C, Mollard P, Garber SS (1995) Modulation of a volume-regulated chloride
current by f-actin. J Membr Biol 147: 283-94.
[7] Abraham R, Weiss A (2004) Timeline - jurkat t cells and development of the t-cell receptor signalling
paradigm. Nature Reviews Immunology 4: 301 -- 308.
[8] Cheng-Mayer C, Iannello P, Shaw K, Luciw PA, Levy JA (1989) Differential effects of nef on hiv
replication: implications for viral pathogenesis in the host. Science 246: 1629-32.
[9] Bausch A, Ziemann F, Boulbitch A, Jacobson K, Sackmann E (1998) Local measurements of vis-
coelastic parameters of adherent cell surfaces by magnetic bead microrheometry. Biophysical Journal
75: 2038 -- 2049.
[10] Bausch A, Hellerer U, Essler M, Aepfelbacher M, Sackmann E (2001) Rapid stiffening of integrin
receptor-actin linkages in endothelial cells stimulated with thrombin: A magnetic bead microrheology
study. Biophysical Journal 80: 2649 -- 2657.
[11] Deng L, Trepat X, Butler JP, Millet E, Morgan KG, et al. (2006) Fast and slow dynamics of the
cytoskeleton. Nature Materials 5: 636 -- 640.
[12] Hoh JH, Schoenenberger CA (1994) Surface morphology and mechanical properties of mdck mono-
layers by atomic force microscopy. J Cell Sci 107 ( Pt 5): 1105-14.
[13] Uhde J, Feneberg W, Ter-Oganessian N, Sackmann E, Boulbitch A (2005) Osmotic force-controlled
microrheometry of entangled actin networks. Phys Rev Lett 94: 198102.
18
[14] Wirtz D (2009) Particle-tracking microrheology of living cells: principles and applications. Annu
Rev Biophys 38: 301-26.
[15] Lu YB, Franze K, Seifert G, Steinhauser C, Kirchhoff F, et al. (2006) Viscoelastic properties of
individual glial cells and neurons in the cns. Proc Natl Acad Sci U S A 103: 17759-64.
[16] Langevin P (1908) The theory of brownian movement. Comptes Rendus Hebdomadaires Des Seances
De L Academie Des Sciences 146: 530 -- 533.
[17] Mason, Weitz (1995) Optical measurements of frequency-dependent linear viscoelastic moduli of
complex fluids. Phys Rev Lett 74: 1250-1253.
[18] Tassieri M, Gibson GM, Evans RML, Yao AM, Warren R, et al. (2010) Measuring storage and loss
moduli using optical tweezers: Broadband microrheology. Physical Review E 81: 026308.
[19] Evans RML, Tassieri M, Auhl D, Waigh TA (2009) Direct conversion of rheological compliance
measurements into storage and loss moduli. Phys Rev E Stat Nonlin Soft Matter Phys 80: 012501.
[20] Mason T, Ganesan K, vanZanten J, Wirtz D, Kuo S (1997) Particle tracking microrheology of
complex fluids. Physical Review Letters 79: 3282 -- 3285.
[21] Kulin S, Kishore R, Hubbard JB, Helmerson K (2002) Real-time measurement of spontaneous
antigen-antibody dissociation. Biophys J 83: 1965-73.
[22] Preece D, Warren R, Evans RML, Gibson GM, Padgett MJ, et al. (2011) Optical tweezers: wideband
microrheology. Journal of Optics 13: 044022.
[23] Gibson GM, Leach J, Keen S, Wright AJ, Padgett MJ (2008) Measuring the accuracy of particle
position and force in optical tweezers using high-speed video microscopy. Opt Express 16: 14561-70.
[24] Yanagishima T, Frenkel D, Kotar J, Eiser E (2011) Real-time monitoring of complex moduli from
micro-rheology. J Phys Condens Matter 23: 194118.
[25] Pesce G, De Luca AC, Rusciano G, Netti PA, Fusco S, et al. (2009) Microrheology of complex fluids
using optical tweezers: a comparison with macrorheological measurements. Journal of Optics A-Pure
and Applied Optics 11: 034016.
19
[26] Addas K, Schmidt C, Tang J (2004) Microrheology of solutions of semiflexible biopolymer filaments
using laser tweezers interferometry. Physical Review E 70: 021503.
[27] Wu P, Huang R, Tischer C, Jonas A, Florin EL (2009) Direct measurement of the nonconservative
force field generated by optical tweezers. Phys Rev Lett 103: 108101.
[28] Cantiello H (1997) Role of actin filament organization in cell volume and ion channel regulation.
Journal of Experimental Zoology 279: 425 -- 435.
[29] Rouse PE (1953) A theory of the linear viscoelastic properties of dilute solutions of coiling polymers.
Journal of Chemical Physics 21: 1272 -- 1280.
[30] Ferry JD (1980) Viscoelastic properties of polymers. Wiley, 3d ed edition.
[31] Rubinstein M, Colby R (2003) Polymer Physics (Chemistry). Oxford University Press, USA.
[32] Pollard TD, Cooper JA (1986) Actin and actin-binding proteins. a critical evaluation of mechanisms
and functions. Annu Rev Biochem 55: 987-1035.
[33] Gardel ML, Shin JH, MacKintosh FC, Mahadevan L, Matsudaira P, et al. (2004) Elastic behavior
of cross-linked and bundled actin networks. Science 304: 1301-5.
[34] Koyama T, Oike M, Ito Y (2001) Involvement of rho-kinase and tyrosine kinase in hypotonic stress-
induced atp release in bovine aortic endothelial cells. J Physiol 532: 759-69.
[35] Farge E, Maggs A (1993) Dynamic scattering from semiflexible polymer. Macromolecules 26: 5041 --
5044.
[36] Tassieri M, Evans RML, Barbu-Tudoran L, Khaname GN, Trinick J, et al. (2008) Dynamics of
semiflexible polymer solutions in the highly entangled regime. Physical Review Letters 101: 198301.
[37] Evans RML (2009) Transforming from time to frequency without artefacts. British Society of Rhe-
ology Bulletin 50: 76.
|
1011.4688 | 2 | 1011 | 2011-08-29T15:35:12 | Rather than resonance, flapping wing flyers may play on aerodynamics to improve performance | [
"physics.bio-ph",
"cond-mat.soft",
"physics.flu-dyn"
] | Saving energy and enhancing performance are secular preoccupations shared by both nature and human beings. In animal locomotion, flapping flyers or swimmers rely on the flexibility of their wings or body to passively increase their efficiency using an appropriate cycle of storing and releasing elastic energy. Despite the convergence of many observations pointing out this feature, the underlying mechanisms explaining how the elastic nature of the wings is related to propulsive efficiency remain unclear. Here we use an experiment with a self-propelled simplified insect model allowing to show how wing compliance governs the performance of flapping flyers. Reducing the description of the flapping wing to a forced oscillator model, we pinpoint different nonlinear effects that can account for the observed behavior ---in particular a set of cubic nonlinearities coming from the clamped-free beam equation used to model the wing and a quadratic damping term representing the fluid drag associated to the fast flapping motion. In contrast to what has been repeatedly suggested in the literature, we show that flapping flyers optimize their performance not by especially looking for resonance to achieve larger flapping amplitudes with less effort, but by tuning the temporal evolution of the wing shape (i.e. the phase dynamics in the oscillator model) to optimize the aerodynamics. | physics.bio-ph | physics | Behind the performance of flapping flyers
Sophie Ramananarivo, Ramiro Godoy-Diana & Benjamin Thiria
Physique et Mécanique des Milieux Hétérogènes (PMMH)
UMR7636 CNRS; ESPCI; UPMC; Université Denis Diderot
10, rue Vauquelin, F-75231 Paris Cedex 5, France
1
1
0
2
g
u
A
9
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
8
8
6
4
.
1
1
0
1
:
v
i
X
r
a
Saving energy and enhancing performance are secular preoccupations shared by both nature and
human beings. In animal locomotion, flapping flyers or swimmers rely on the flexibility of their wings
or body to passively increase their efficiency using an appropriate cycle of storing and releasing elastic
energy. Despite the convergence of many observations pointing out this feature, the underlying
mechanisms explaining how the elastic nature of the wings is related to propulsive efficiency remain
unclear. Here we use an experiment with a self-propelled simplified insect model allowing to show
how wing compliance governs the performance of flapping flyers. Reducing the description of the
flapping wing to a forced oscillator model, we pinpoint different nonlinear effects that can account
for the observed behavior -- in particular a set of cubic nonlinearities coming from the clamped-
free beam equation used to model the wing and a quadratic damping term representing the fluid
drag associated to the fast flapping motion. In contrast to what has been repeatedly suggested in
the literature, we show that flapping flyers optimize their performance not by especially looking for
resonance to achieve larger flapping amplitudes with less effort, but by tuning the temporal evolution
of the wing shape (i.e. the phase dynamics in the oscillator model) to optimize the aerodynamics.
INTRODUCTION
Flying animals have since long inspired admiration and
fueled the imagination of scientists and engineers. Along-
side biologists studying form and function of flapping fly-
ers in nature [1, 2], the last decade has seen an impressive
quantity of studies driven by engineering groups using
new techniques to develop and study artificial biomimetic
flapping flyers [3, 4]. The widespread availability of high-
speed video and in particular the merging of experimental
methods borrowed from fluid mechanics into the toolbox
of the experimental biologist have permitted to elucidate
various key mechanisms involved in the complex dynam-
ics of flapping flight (see e.g. [5 -- 7]).
A recent field of investigation concerns the efficiency of
flapping flyers, the major interrogation being about how
natural systems optimize energy saving together with
performance enhancement. In particular, the passive role
of wing flexibility to increase flight efficiency through the
bending of the wings while flapping has attracted a lot of
attention. It is commonly agreed that this efficiency en-
hancement comes from the particular shape of the bent
wing, which leads to a more favorable repartition of the
aerodynamic forces (see [8] and [9] for an extensive re-
view). For flying animals in air, such as insects, it has
been proposed [10 -- 12] that wing inertia should play a
major role in competing with the elastic restoring force,
compared to the fluid loading. The mechanism govern-
ing the propulsive performance of the flapping flyer can
therefore be seen at leading order as a two-step process,
where the instantaneous shape of the wings is determined
by a structural mechanics problem which then sets the
moving boundaries for the aerodynamic problem.
From a dynamical point of view, if we consider chord-
wise bending of a wing with a given flapping signal im-
posed at the leading edge, the instantaneous shape of
the structure is strongly dependent on the phase lag be-
tween the forcing and the response of the wing (respec-
tively the leading and trailing edges). Recent works by
[13] and [14] using a simplified model of a flexible wing
as a combination of heaving and passive pitching have
shown that a transition from enhanced thrust to under-
performance occurs for a critical phase value close to the
resonant frequency of the system. This sustains the com-
monly invoked argument suggesting that flapping flyers
could take advantage of a structural property to save en-
ergy by matching the relaxation frequency of their com-
pliant wings to the wingbeat frequency [13, 15 -- 17]. In
nature this has been observed in particular for undu-
latory swimming fish or other swimmers that use de-
forming propulsive structures, such as jellyfish or scal-
lops (see [18] and references therein). In the case of in-
sects, however, the few available observations (especially
for large species) report wingbeat frequencies far below
the natural relaxation frequencies [19 -- 22]. Recent exper-
iments using a self-propelled model with large-flapping-
amplitude elastic wings [12] are consistent with the lat-
ter, since the propulsive efficiency of the model peaks
for a flapping frequency lower than the primary linear
resonance of the wings. Fully predicting the wing beat
rate as the undamped resonant frequency of a linear os-
cillator (see e.g.
[15]) should be therefore taken with
reserve. Super-harmonic nonlinear resonances have been
invoked [23], suggesting that flying animals may effec-
tively flap their wings far below the primary resonance
while increasing their performance. This is probably one
mechanism among others governing the dynamics of flap-
ping flyers, but it is clear that the details of the underly-
ing fluid-structure interaction problem are poorly under-
stood. More specifically, the underlying phase dynamics
2
EXPERIMENTS
Setup and physical quantities
The experimental setup is the same described by Thiria
& Godoy-Diana [12], inspired from the pioneer 19th cen-
tury experiment by Marey [24]: a flapping wing device is
attached to a mast that is ball bearing mounted to a cen-
tral shaft in such a way that the thrust force produced by
the wings makes the flyer turn around this shaft. A par-
ticular attention has been paid to reduce friction losses
in the whole system. Wings are made of Mylar R(cid:13) semi-
circles of diameter S = 2L = 6 cm. The experimental
parameters are the forcing frequency (f), the flapping
amplitude (Aω) and the chordwise rigidity of the wings
(B) governed by their thickness h. In contrast with the
first study reported with this setup [12], the set of wings
used here covers a larger range of bending rigidities, from
near-rigid to very soft materials. Six pairs of wings have
been tested. Their structural properties (thickness, mass,
and rigidity) are summarized in Table 1.
This specific setup allows to measure various averaged
quantities (see [12] for details):
the cruising speed U
when the device is allowed to turn around, and the thrust
force FT when it is held at a fixed station (see Fig. 2 (a)
and (b)) which gives the averaged aerodynamic thrust
power, being the product PT = U F . In both cases, the
power consumption Pi is measured. On the other hand,
we performed a precise dynamical study of the flapping
wing. For each set of parameters (Aω, ff , B), the phase
and amplitude of the trailing edge, with respect to the
forcing flapping motion, has been measured using a fast
cadenced camera (1000 fps) in both air and vacuum.
It is important to recall that for this setup, and more
generally for flapping flyers in air, the main bending mo-
tor of the flexible wings is wing inertia [10 -- 12]. The com-
petition between the wing inertia and the elastic restoring
force is captured by the scaled elasto-inertial number Nei
[12]:
(cid:18) ωf
(cid:19)2
(1)
Nei =
µsAwω2
f L3
B
=
Aω
L
ω0
The first expression is a direct comparison between
both the moments of inertial and elastic forces.
Inter-
estingly, this number can also be expressed as a func-
tion of the ratio between the forcing and relaxation fre-
quencies times the non-dimensional forcing amplitude of
the driving motion, which allows to express directly the
bending rate as function of a non-dimensional oscillator
forcing term. The second expression is therefore useful
to explore the nearness of the resonance and will be used
to analyze the experimental data in this paper. Results
will therefore displayed as a function of the reduced fre-
quency ¯ωf = (ωf /ω0) = ¯Aω−1/2N 1/2
L is
, where ¯Aω = Aw
ei
FIG. 1: Experimental setup: a) Pioneer experiment from
Marey [24]. b) Actual setup. c) Details of the flapping flyer
model used for this study.
that set the instantaneous wing shape and lead first to
an increase and then a loss of the thrust power (and even
a reversal of the propulsive force as in the case of [13])
remain unexplained.
In this paper we address these questions using the ex-
perimental self-propelled flapping-wing model with elas-
tic wings described in [12]. Exploring a wide range of
bending rigidities we show that, in the simplified context
of chordwise-compliant wings, the performance optima of
the system are far from being set by a simple resonant
condition. We develop a nonlinear one-dimensional beam
model for the bending wing which is reduced to a forced
oscillator model suitable to study different nonlinear ef-
fects. In particular, a set of cubic nonlinearities coming
from the clamped-free beam equation and a quadratic
damping term representing the fluid drag associated to
the fast flapping motion permit to account for the ob-
served behavior. We show that the nonlinear nature of
the fluid damping is an essential feature to determine
the phase lag that leads to an increase/decrease of the
efficiency.
As a whole fluid-solid interaction process leading to
propulsion, we provide evidence that flapping flyers may
optimize their performance not by especially looking for
resonance but by using passive deformation to streamline
the instantaneous shape of the wing with the surrounding
flow.
(cid:1)(cid:2)(cid:3)(cid:1)(cid:4)(cid:1)(cid:5)(cid:6)(cid:7)(cid:8)(cid:6)(cid:5)(cid:9)(cid:5)(cid:10)(cid:6)(cid:11)(cid:7)(cid:9)(cid:12)(cid:10)(cid:3)(cid:13)(cid:6)(cid:5)(cid:6)(cid:8)(cid:14)(cid:15)(cid:9)(cid:8)(cid:3)(cid:2)(cid:6)(cid:11)(cid:11)(cid:15)(cid:2)(cid:5)(cid:10)(cid:11)(cid:14)(cid:8)(cid:6)(cid:16)(cid:3)(a)(b)(c)3
FIG. 2: (a) Cruising speed, (b) thrust force and nondimen-
sional (c) thrust (pT ) and (d) input (pi) powers as a function
of ¯ωf . The gray area represents the optimum region, the
dashed line indicates the location of the reduced natural fre-
quency of the wing (linear resonance).
the reduced flapping amplitude. In order to compare the
aerodynamic performance in all the experiments, both
the thrust force and cruising speed were rendered non-
dimensional using the appropriate scalings fT = FT L/B
and u = U/Aωω. The non-dimensional powers (displayed
in Fig. 2 (c) and (d)) then read pT = U FT L/BAωω and
pi = PiL/Bω.
In both the thrust force and cruising speed curves, it
is clear that increasing wing flexibility brings out two
distinct regimes: up to a certain flapping frequency, the
more flexible wings outperform the rigid linear U (f ) re-
lationship (see also [25]). The measurements for the two
most flexible wings evidence the appearance of an under-
performance regime in which both FT and U lie below
the rigid wing case. Looking now at the nondimensional
thrust power, the data from all wings collapse on a single
curve with a clear performance peak, which agrees with
what has been observed by [13, 14] for heaving/pitching
systems. An important point is that the maximum in
performance does not take place at the resonant fre-
quency, but much below (around 0.7ω0, represented by
the gray shaded area). Moreover, the nondimensional
thrust power at ¯ωf = 1 (see dahsed line in Fig. 2 (c)) is
even more than 4 times lower than the optimum value.
At last, we remark that there is also no sign of a resonant
behavior in the consumed power curve (Fig. 2 (d)).
Wing dynamics
We proceed now to study the behavior of the wings
considered as a forced oscillator, assuming the oscillation
FIG. 3: a): Photograph of the flapping wing showing succes-
sive states of the bending wing during one stroke cycle (thick-
ness is 0.050 mm and ¯ωf = .5). As can be seen, the main
deformation is mainly performed on the first mode. In this
case the phase lag is quite large, leading to a strong increase
of flight performance. b): Typical time series tracking the
motion of the leading (black curve) and trailing (red curve)
edges of the wing at mid-span, obtained from video record-
ings at 1000 fps. c): Same as b) but with a forcing near 1
3 ω0,
exhibiting super-harmonic resonance typical from dynamical
systems containing cubic nonlinearities.
of the leading edge to be the forcing and that of the
trailing edge to be the response (which means to assume
that the wings bend following only the first deformation
mode). As said before, the amplitude and phase shift
of the response can thus be measured by following the
two wing edges on a high cadenced camera recording (as
seen on Fig. 3 (a)). Figs. 3 (b) and (c) display two
characteristic time evolutions of the driving oscillation
(the imposed wing beat, shown as black dots) and the
wing elastic response (the motion of the trailing edge,
red dots) in the moving frame. The first case shows a
typical response, at ¯ωf = 0.79, mainly sinusoidal at the
driving frequency, which supports the assumption that
the oscillations of the wing follow a single mode. In the
second case, the driving frequency is near one third of
the resonant frequency ω0. As can be observed in Fig. 3
(c), the response is then a combination between ω0/3 and
ω0, giving evidence of a super-harmonic resonance [26],
pointing out the fact that the system integrates cubic
nonlinearities. The non dimensional amplitude a (i.e.
scaled by the length of the wing L) and phase γ have
therefore been extracted from those signals for each pair
of wings as a function of the reduced driving frequency for
two different amplitudes. Results are displayed in Fig. 4.
In parallel, the same experiments have been conducted
in a vacuum chamber at 10 % of the ambient pressure.
Results are also displayed in Fig. 4 for comparison.
10203000.010.020.030.04f(Hz)FT(N)1015202530350123f(Hz)U(m.s−1)00.5100.511.52¯ωfpt00.510200400600800¯ωfpi(b)(c)(d)nearrigidcase(a)00.020.040.060.08 0.1 0.020.04 t(s)(b)(c)(a)Table 1. Wing properties
wing thickness, h (mm)
mass per unit area µs (kg.m−2)
rigidity B (N.m)
relaxation frequency f0 (Hz)
color label in figures
0.050
4.50 10−2
3.34.10−5
25.4
blue
0.078
10.63 10−2
1.83.10−4
34.2
red
0.130
17.67 10−2
1.02.10−3
62.2
green
0.175
24.12 10−2
2.26.10−3
89.5
yellow
0.250
34.92 10−2
7.31.10−3
117.1
purple
0.360
47.95 10−2
14.00.10−3
160.8
black
4
As can be seen, the evolution of the amplitude a shows
a fast increase from very low flapping frequencies. This
is the expected behavior owing to the inertial charac-
ter of the forcing. A slight but rather broad peak can
be observed in the nearness of ω0/3 in the amplitude
curve, confirming the occurrence of the super-harmonic
resonance hinted above and strengthening the fact that
this type of mechanism may play a role as a strategy
for performance enhancement in nature [23]. Two more
points have to be underlined: first, measurements in air
and vacuum are approximately the same, in accordance
with the hypothesis that inertia is the main bending fac-
tor for flapping flyers [10 -- 12]. The second point is that
no clear resonance is observed around ¯ωf = 1 (only a
barely visible peak in the case of the lowest forcing am-
plitude shown in the insert in Fig. 4(a)). Concerning the
phase γ, the present results recover the trend of what
γ increases
has been observed recently [9, 13, 14, 16]:
monotonically with ¯ωf . Considering the experiments in
air at normal conditions, this observation together with
the performance increase shown in the first part of the
pt(¯ωf ) (Fig. 2 (c)), brings the following conclusion: the
increasing phase shift γ, which corresponds to a situation
where the wing experiences a larger bending at the max-
imal flapping velocity, leads to a more favorable reparti-
tion of the aerodynamic forces (as discussed in [12]).
A simple argument widely shared in the community con-
necting the phase dynamics to the propulsive perfor-
mance is: the larger the phase lag is, the best the thrust
power would be [13, 14], until the point where the wing
experiences its largest bending at γ = π/2. However,
while the argument reasonably agrees with the observa-
tions in the range of forcing frequencies where perfor-
mance increases with ¯ωf , the maximum performance does
not actually match with the maximum of bending that
occurs at γ = π/2, but relatively far below this expected
optimum (which lies actually around π/4).
One last important remark to be made concerns the
phase evolution in vacuum.
It is clearly observed that
γ decreases more slowly in the low density environment
within the whole range of flapping frequencies studied.
In contrast with the amplitude measurements, where
the data from the experiments in vacuum follow roughly
the same curve of those in air at atmospheric pressure,
the large difference in the γ curves between both cases
points out unequivocally the importance of the surround-
ing fluid in determining the phase dynamics. This point
will be discussed later. At this stage, we have shown that,
FIG. 4: Evolution of the non-dimensional amplitude a) and
phase b) of the trailing edge wing response as a function of the
reduced driving frequency for both flapping amplitudes ¯Aω =
0.8 and ¯Aω = 0.5 (filled symbols correspond to measurements
in air, open symbols in vacuum). Those results are compared
to nonlinear predictions from Eq. 12 with (gray line) and
without (black line) nonlinear air drag (discussed further in
the text).
as observed in the pitching/heaving systems of [13, 14],
the increase in performance of elastic wings undergoing
large oscillations is essentially governed by a fast grow-
ing phase evolution. However, the physical mechanisms
governing the propulsive performance remain unclear. In
particular, the mechanisms leading to the useful evolu-
tion of γ as well as the link between resonance and per-
formance are still looking for a definitive answer.
NONLINEAR 1D BEAM MODEL
In order to understand those crucial points, one can
consider the elastic wing as a clamped-free beam under
base harmonic forcing. For simplicity, the beam is con-
sidered as one-dimensional taken at mid length in the
spanwise direction of the wing. We assume here, accord-
ing to the experiment, that only flexural displacements
(i.e. perpendicular to the direction of the flight motion)
are allowed. The structural properties of the beam are
determined by measuring experimentally the relaxation
frequency.
Thus, the equation governing the motion of the non-
linear flexural oscillations of clamped-free beam writes
[27]:
00.511.500.20.40.60.81a¯ωf0.511.50.20.40.60.800.511.5(cid:239)pi/2(cid:239)pi/40γ(rad)¯ωf(a)(b)EIW (cid:48)(cid:48)(cid:48)(cid:48) + µ W = −EI(W (cid:48)W (cid:48)(cid:48)2 + W (cid:48)(cid:48)(cid:48)W (cid:48)2)(cid:48)
(cid:34)
(cid:90) x
L
− µ
2
W (cid:48)
∂2
∂t2
(cid:35)
(cid:35)(cid:48)
(2)
W (cid:48)2dx
dx
(cid:34)(cid:90) x
0
where W is the transversal local displacement, E the
Young modulus, I the second moment of inertia and µ
the mass per unit of length. Writing W as W (x, t) =
w(x, t) + w0(t), where w0(t) is the driving motion defined
by w0(t) = Aω cos(ωf t), and using the non-dimensional
τ ; with
quantities for space and time w = w
L ; x = x
L; t = t
τ =(cid:0) µ
EI
(cid:1)1/2
L2, equation 2 reads:
w(cid:48)(cid:48)(cid:48)(cid:48) + w = −( w(cid:48) w(cid:48)(cid:48)2 + w(cid:48)(cid:48)(cid:48) w(cid:48)2)(cid:48)
(cid:34)
(cid:90) x
1
− 1
2
w(cid:48)
∂2
∂t2
(cid:34)(cid:90) x
0
(cid:35)
(cid:35)(cid:48) − ¯Aω w0(3)
w(cid:48)2dx
dx
which has to satisfy the clamped-free boundary condi-
tions w(0, t) = w(cid:48)(0, t) = w(cid:48)(cid:48)(1, t) = w(cid:48)(cid:48)(cid:48)(1, t) = 0. The
last term on the right hand side in Eq. 3, − ¯Aω w0 =
2 cos( ¯ωf t) = Nei cos( ¯ωf t), is a forcing term due to
¯Aω ¯ωf
the wing inertia whose amplitude is given by the elasto-
inertial number and which is dependent on the square of
the driving frequency as seen before.
The next step is to set apart the spatial dependence by
projection of Eq. 3 onto the complete set of eigenfunc-
tons defined by the linear part. The displacement is ex-
pended as w(x, t) =(cid:80)∞1 Xp(t)Φp(x) (see [28]) where Φp
are the non-dimensional linear modes for clamped-free
beams which are not recalled here for the sake of brevity.
The problem then writes (the have been removed for
simplicity):
i,j,k=1
(cid:90) 1
(cid:90) x
0
Φ(cid:48)i
hp
ijk =
(Φ(cid:48)iΦ(cid:48)(cid:48)j Φ(cid:48)(cid:48)k + Φ(cid:48)(cid:48)(cid:48)i Φ(cid:48)jΦ(cid:48)k)(cid:48)Φpdx
(5)
(cid:34)
(cid:90) 1
0
f p
ijk =
(cid:90) u
(cid:35)(cid:48)
Φ(cid:48)j(y)Φ(cid:48)k(y)dydu
Φpdx
(6)
1
0
The projection of the forcing term on the pth mode,
Fp, writes at the trailing edge:
Xp + Xp = − N(cid:88)
− N(cid:88)
i,j,k=1
hp
ijkXiXjXk
where hp
ijk and f p
ijk are determined by:
ijk(XiXj Xk + Xi Xj Xk) + Fp(t)(4)
f p
(cid:90) 1
0
Φp(x)dx
Fp = ¯Aω ¯ωf
2Φp(1)
5
(7)
As the propulsive regimes observed in this work lie be-
low the first relaxation frequency of the wing, we assume
that the response of the wing is mainly governed by the
first eigenmode. Hence, equation 4 can be considerably
simplified and reduces for the only mode 1 to:
X + X = −h1
111X 3 − f 1
111(X 2 X + X X 2) + F1(t)
(8)
A crucial feature is now to choose a damping term
to this dynamical system. During a stroke cycle, the
wing follows very fast motions involving high local
Reynolds numbers, which prompt us to include a non-
linear quadratic fluid drag term [29] in addition to the
classical linear viscous friction law. The damping is then
chosen as a combination of linear and nonlinear terms as
follows:
Ξ(X,
X) = ξ X + ξnl X X
(9)
The linear and nonlinear coefficients ξ and ξnl are
estimated studying the impulse response for each wing
[26].The solution of Eq. 8 including damping is deter-
mined by using a classical multiple scale method at first
order (see [26]). To this end, we introduce a small pa-
rameter and a detuning parameter σ = ( ¯ωf − 1)/. The
problem to be solved reads.
X + X = −(h1
+ Ξ(X,
111X 3 + f 1
X) + F1(t))
111(X 2 X + X X 2)
(10)
According to the multiple scales theory, we express
the solution in terms of different time scales as X =
X0(t0, t1) + X1(t0, t1) + .... where t0 = t and t1 = t
are respectively short (relative to the oscillation of the
wing) and long times scales. The system at order 0 is
t0X0 + X0 = 0 an gives the straightforward solution
∂2
X0 = A(t1)eit0 + A∗(t1)e−it0 where A and A∗ are com-
plex functions.
At order 1, we obtain:
t0 X1 + X1 = −h1
∂2
− Ξ(X0,
111X 3
111(X 2
0
X0 + X0 X0
0 − f 1
X0) − 2∂t1t0X0 + F1 cos(t0 + σt1)
(11)
)
2
Using the expression of X0 found at order 0 into Eq.
11, an equation for A is obtained by elimination of the
secular terms:
AA) =
4ξnl
3π
1
2
111−2f 1
A2A∗(3h1
111)+i(2∂t1 A+ξA+
F1eiσt1
(12)
where the pre-factor 4
3π in front of the nonlinear damp-
ing coefficient is obtained during the special integration
over one period of the Fourier expansion of the function
X0 X0 (see [26]).As can be seen, Eq. 12 is a charac-
teristic equation of a forced damped oscillator with cu-
bic nonlinearities. At last, substituting the polar form
2 aei(σt1−γ), separating into real and imaginary parts
A = 1
and looking only to the steady-state solutions, we find
two relations for the amplitude a and phase γ.[32]
(cid:0)Γ1a3 − aσ(cid:1)2
+ (ξa +
ξnla2)2 =
(cid:18) (ξa + 4
4
3π
Γ1a3 − aσ
F 2
1
4
3π ξnla2)
(cid:19)
γ = arctan
(13)
(14)
111 − 2f 1
where Γ1 = 1
coefficient, which is computed from Eq. 5 and 6.
111) is the nonlinear cubic term
8 (3h1
Eq. 14 closely resembles a classic nonlinear Duffing
oscillator except that the forcing amplitude is frequency
dependent and that a nonlinear damping term is present.
ANALYSIS AND DISCUSSION
Resonance and phase evolution
Predictions of the above model for the parameters of
the experiments are plotted in Fig. 4 for both cases in
air and vacuum. In addition, for a clear understanding
of the underlying dynamics described by Eqs. 13 and 14,
a comparison between predictions from a linear model,
a nonlinear with linear damping and a nonlinear with
nonlinear damping is displayed in Fig. 5 for two flapping
amplitudes ¯Aω. It can be seen that the model based on a
single mode is capable of reproducing all the observations
made from the experiments both in normal and low den-
sity environments. The good agreement between experi-
ments and model allows us to pinpoint some mechanisms
underlying the complex mechanisms of flapping flight.
The first concerns the question of resonance: from Fig. 5,
it can be observed that the only case (apart from the lin-
ear case) exhibiting a slight resonance peak corresponds
to relatively small flapping amplitude and damping co-
efficient [i.e. only linear damping term, see Fig. 5 (a)].
Cases for higher amplitude and/or presence of nonlin-
ear damping behave as a non-resonant like system in the
range of flapping frequencies studied. In nonlinear oscil-
lators, it is known that the main effect of the nonlinear
term is to distort the resonance curve and shift the reso-
nance peak to higher frequencies (for a hardening coeffi-
cient Γ1 > 0, as in the present study) [26]. An important
6
FIG. 5: Dependence of the amplitude a and phase γ with
the reduced forcing frequency ¯ωf for the first mode of a
clamped-free beam forced by inertia for two different (high
and medium) amplitudes Aω (chosen arbitrarily for clarity).
The blue line corresponds to the linear prediction, the black
line to the non-linear model from Eq 12 with linear damping
(ξnl = 0), the red line to the non-linear model with nonlinear
damping (ξnl (cid:54)= 0). As can be seen only cases with relatively
small flapping amplitude and linear damping can exhibit a
slight resonance peak. Greater amplitudes and/or presence
of nonlinear damping behave as a non-resonant system in the
domain of flyers capabilities. Concerning the phase, models
including only linear damping do not produce "useful" phase
lag except in the nearness of the phase jump. In contrast, the
presence of a nonlinear damping produces a fast and helpful
evolution.
feature of such nonlinear systems is that the distortion of
the shape of the resonance curve is directly dependent on
the amplitude of the excitation. In the present case where
the forcing is inertial, the response depends on the square
of the forcing frequency (or on the elasto-inertial number
Nei), which provides an increase of the amplitude plotted
in Fig. 4 independent of an intrinsic resonance mecha-
nism. Hence, we can expect the actual resonance curve
of the system to be all the more distorted that the flap-
ping frequency increases. Another feature that makes it
difficult for the flapping flyer to benefit from a resonance
mechanism is the presence of a geometric saturation due
to the finite length of the wing. Always due to the in-
ertia effects, this geometrical saturation will be reached
all the more soon that the demand for larger amplitude
(i.e. better performances) is increased. Coming back to
the distorted resonance curve, the visible consequence is
that the wing, even for a small nonlinear cubic coefficient,
behaves as a system never reaching a peak in the range
of frequencies commonly used by flapping flyers. Addi-
tionally, the presence of strong damping accentuates this
behavior by smoothing the value of a possible resonance
peak. This last observation is consistent with the fact
that birds or insects may not especially look for struc-
01230123a0123012345a0123(cid:239)3(cid:239)2(cid:239)10¯ωfγ0123(cid:239)3(cid:239)2(cid:239)10¯ωfγAω/L=0.45Aω/L=1.2ξnl=0.4ξnl=0tural resonance to improve their performance.
The second point is the crucial role of fluid damping
in triggering the phase lag that is useful for thrust en-
hancement. For the phase, shifting the resonance peak
as a result of the nonlinear spring in the oscillator model
means shifting the phase jump at γ = π/2 to higher
frequencies as well. Thus, without air drag, as can be
seen in Fig. 5 (c) and (d), the nonlinear evolution of
the phase γ(¯ωf ) would be even slower than in the linear
case for which the phase evolution is already not espe-
cially favorable except in the nearness of the resonance.
This is exactly what is observed for the vacuum measure-
ments where the nonlinear damping due to fluid drag is
negligible. On the contrary, the presence of a quadratic
fluid damping determines a fast increase of the phase lag
(and a so a thrust improvement) even from the very first
flapping frequencies. This implies of course that strong
flapping velocities are a necessary condition for the bend-
ing to become efficient (i.e. elasticity will play a minor
role if the flapping beat amplitude is not strong enough).
Summarizing, the instantaneous wing shape is given by
the two following ingredients: inertia provokes the bend-
ing (gives the amplitude) and damping, by controlling the
phase lag, allows this bending to be usefully exploited.
Large phase lags will provide largest bending of the wing
at maximum flapping speed, leading to a more favorable
repartition of aerodynamics forces.
Optimum
Since classic resonance mechanisms cannot answer it,
the question of the performance optimum (or the tran-
sition to underperformance) remains unclear. We there-
fore proceeded to study the kinematics of the wing in
the laboratory frame. In particular, we have compared
both characteristic angles relative to the global wing
motion. The first characteristic angle is dependent on
the ratio between the maximal vertical flapping veloc-
ity uω = ωAω and the cruising velocity U and reads:
φ = arctan(ωAω/U ). φ is considered as the instanta-
neous angle of attack of the wing and as can be seen,
is directly related to the Strouhal number St = ωAω/U
which determines as well the performance of flapping fly-
ers [30]. We define a second characteristic angle θ as
the trailing-edge angle taken at the maximum flapping
velocity. This angle is directly related to the phase lag
γ, and thus determines to what extent the bending of
the wing will be useful in terms of performance. Fig. 6
shows the evolution of the ratio θ/φ, which is naturally a
growing function of ¯ωf because both an increase in θ or
a decrease in φ lead to an enhancement of the propulsive
performance.
The interesting point is that the location of the per-
formances/under performances transition takes place at
θ/φ = 1 (i.e. when both angles point instantaneously at
7
FIG. 6: Evolution of the two characteristic angles of the wing
motion θ and φ as a function of the reduced driving frequency
¯ωf . Two regimes can be distinguished: (I): φ < θ correspond-
ing to the performances increasing stage due to a useful phase
lag.
(II): φ > θ corresponding to the transition to under-
performances due to a loss of the effective wing area. The
optimum occurs therefore when φ and θ point at the same
direction (best phase lag).
the same direction). Thus, the optimum value of θ does
not corresponds to the maximum bending experienced by
the wing (which would be the optimal solution) but to
the moment when the deflection angle matches the angle
of attack as sketched in Fig. 6. For a rigid wing, because
θ is fixed (= 0), the optimization problem is here nonex-
istent and thrust only depends on the driving frequency
(for a given amplitude). With flexibility and according
to what has been previously observed, θ starts increasing
and tends to align the wing trailing-edge with the flow.
As discussed earlier, this leads to a more favorable repar-
tition of the aerodynamics forces as sketched in Fig. 6.
However, this argument is only valid if the surrounding
flow is totally attached to the wing (i.e. separation oc-
curs only at the trailing-edge). A situation where θ > φ
is strongly subjected to flow separation before the wing
trailing edge. In this case the effective surface relative to
the aerodynamic load can be expected to be drastically
reduced leading to a loss of aerodynamic performance. It
has to be noticed that the value of π/2, or more gener-
ally values of phases greater than θopt observed in this
experiments should be, theoretically, more optimal (i.e.
should give more optimal bending shapes for useful pro-
jection of forces). However, if a separation occurs, the
corresponding loss of thrust force (and so cruising speed)
will accelerate the decoherence of both angles and hence,
will provoke the subsidence of the performance, as has
been observed on Fig. 2. The more economic strategy
to fly is therefore to set θ ≈ φ which corresponds to the
optimum way to transfer useful momentum.
0.40.60.811.200.511.5¯ωfθ/φ(II)(I)CONCLUDING REMARKS
In this work, we aimed at describing the dynamics gov-
erning the performance of flapping flyers. Considering
large flapping amplitude and relatively large wings (as for
big insect species), we have shown that nonlinear and in-
ertia effects, together with geometric limitation, question
the prevailing idea that energy-saving strategies in flap-
ping flight must be related to resonance mechanisms. In
search of improving performances, animals may actually
stay below the resonance point. Besides, the nonlinear
nature of air drag (which implies sufficiently strong flap-
ping amplitudes) seems to be a fundamental ingredient
to create the phase lag between the leading and trailing
edges of the flapping wing that allows the elasticity en-
ergy to be used at its best. One last comment is that
the presence of structure resonances for flyers in nature
is not invalidated by the mechanism described here. For
instance, small insects may not use much elasticity and
bending because either their wings are too small or the
local Reynolds number is not sufficiently high to produce
enough damping, and thus a useful phase lag. However,
studies containing a large bank of comparative resonant
frequencies and wingbeats of insects or birds being rare in
the literature, it is consequently hard to draw any conclu-
sion about the existence of two distinct strategies at this
state. According to biologists, resonant mechanisms lie
at the muscle level more than in the wing structure itself
(see [2, 31] and reference therein) which would strengthen
that there is no reason, a priori, for flapping flyers to look
for structural resonance of the wing. Further analysis on
such a way would certainly help to discern if there are,
or not, universal characteristics for flapping flyers.
The authors are grateful to Daniel Pradal for his help
concerning the experimental setup, Cyril Touzé for hav-
ing shared his knowledge of nonlinear systems and Sarah
Tardy for her careful reading of the manuscript. This
work was supported by the French Research Agency
through project ANR-08-BLAN-0099.
[1] David E. Alexander. Nature's Flyers: Birds, Insects, and
the Biomechanics of Flight. The Johns Hopkins Univer-
sity Press, 2004.
[2] R. Dudley. The Biomechanics of Insect Flight. Princeton
University Press, 2000.
[3] S Ho, H Nassef, N Pornsinsirirak, YC Tai, and CM Ho.
Unsteady aerodynamics and flow control for flapping
wing flyers. Progress in Aerospace Sciences, 39(8):635 --
681, 2003.
[4] Wei Shyy, Yongsheng Lian, Jian Tang, Dragos Viieru,
and Hao Liu. Aerodynamics of low Reynolds number fly-
ers. Cambridge Aerospace Series. Cambridge University
Press, 2008.
[5] MH Dickinson, FO Lehmann, and SP Sane. Wing rota-
tion and the aerodynamic basis of insect flight. Science,
8
[6] ZJ Wang. Dissecting insect flight. Annu. Rev. Fluid
284(5422):1954 -- 1960, 1999.
Mech., 37:183 -- 210, 2005.
[7] Geoffrey R Spedding and Anders Hedenström. Piv-based
investigations of animal flight. Exp Fluids, 46(5):749 -- 763,
2009.
[8] J. M. Anderson, K. Streitlien, D. S. Barret, and M. S. Tri-
antafyllou. Oscillating foils of high propulsive efficiency.
J. Fluid Mech., 360:41 -- 72, 1998.
[9] W. Shyy, H. Aono, S.K. Chimakurthi, P. Trizila, C.-K.
Kang, C.E.S. Cesnik, and H. Liu. Recent progress in
flapping wing aerodynamics and aeroelasticity. Progr.
Aerospace Sci., 2010.
In press, corrected proof. DOI:
10.1016/j.paerosci.2010.01.001.
[10] T. L. Daniel and S. A. Combes. Flexible Wings and Fins:
Integr.
Bending by Inertial or Fluid-Dynamic Forces?
Comp. Biol., 42(5):1044 -- 1049, 2002.
[11] S.A. Combes and T.L. Daniel. Into thin air: contribu-
tions of aerodynamic and inertial-elastic forces to wing
bending in the hawkmoth manduca sexta. J. Exp. Biol.,
206:2999 -- 3006, 2003.
[12] B. Thiria and R. Godoy-Diana. How wing compliance
drives the efficiency of self-propelled flapping flyers. Phys.
Rev. E, 82:015303(R), 2010.
[13] S. E. Spagnolie, L. Moret, M. J. Shelley, and J. Zhang.
Surprising behaviors in flapping locomotion with passive
pitching. Phys. Fluids, 22(4):041903, 2010.
[14] J. Zhang, L. Nan-Sheng, and L. Xi-Yun. Locomotion of
a passively flapping flat plate. j. fluid mech. in press. J.
Fluid Mech., 659:43 -- 68, 2010.
[15] C.H. Greenewalt. The wings of insects and birds as me-
chanical oscillators. Proc. Nat. Acad. Sci., 104:605 -- 611,
1960.
[16] H. Masoud and A. Alexeev. Resonance of flexible wings
at low Reynolds number. Phys. Rev. E, 81:056304, 2010.
[17] S. Michelin and S. G. Llewellyn Smith. Resonance and
propulsion performance of a heaving flexible wing. Phys.
Fluids, 21(7):071902, 2009.
[18] J. H. Long and K. S. Nipper. The importance of body
stiffness in undulatory propulsion. Amer. Zool., 36:678 --
694, 1996.
[19] S. Sunada, L. Zeng, and K. Kawachi. The relationship
between dragonfly wing stricture and torsional deforma-
tion. J. Theor. Biol., 193:39 -- 45, 1998.
[20] S. Sunada and et al. Optical measurement of the de-
formation motion, and generated force of the wings of
a moth, mythimna separa (walker). JSME International
Journal Series B, 45:836 -- 842, 2002.
[21] M. Nakamura, A. Iida, and A. Mizuno. Visualization of
three-dimensional vortex structures around a dragonfly
with dynamic piv. J. Visualization, 10:159 -- 160, 2007.
[22] J. S. Chen, J.-Y. Chen, and Y.-F. Chou. On the natu-
ral frequencies and mode shapes of dragonfliy wings. J.
Sound Vib., 313:643 -- 654, 2008.
[23] M. Vanella, T. Fitzgerald, S. Preidikman, E. Balaras,
and B. Balachandran. Influence of flexibility on the aero-
dynamic performance of a hovering wing. J. Exp. Biol.,
212(1):95 -- 105, Jan 2009.
[24] Antoine Magnan. La locomotion chez les animaux: I-Le
vol des insectes. Hermann & Cie., 1934.
[25] N. Vandenberghe, J. Zhang, and S. Childress. Symmetry
breaking leads to forward flapping flight. J. Fluid Mech.,
506:147 -- 155, 2004.
[26] A.H. Nayfeh and D.T. Mook. Nonlinear oscillations.
John Wiley & sons, New-York, 1979.
[27] M.R.M Crespo Da Silva and C.C. Glynn. Nonlin-
ear flexural-flexurale-torsional dynamics of inextensional
beams. II. Forced Motions. J. Struct. Mech., 6:449 -- 461,
1978.
[28] A.H. Nayfeh. Method of Normal Forms. John Wiley &
[30] G. K. Taylor, R. L. Nudds, and A. L. R. Thomas. Flying
and swimming animals cruise at a strouhal number tuned
for high power efficiency. Nature, 425:707 -- 711, 2003.
[31] P. W. Willmott and C. P. Ellington. The mechanics of
[29] D. J. Tritton. Physical Fluid Dynamics. Oxford Univer-
sons, New-York, 1993.
sity Press, 1988.
9
flight in the hawkmoth manduca sexta. J. Exp. Biol.,
200:2705 -- 2722, 1997.
[32] It has to be noted that the linear damping term aξ cor-
responds to structural damping (and viscous fluid damp-
ing relative to very small displacements) and is therefore
mainly dependent on the only displacement X (i.e. in the
3π ξnla2 is strongly dependent
wing frame). In contrast, 4
on the global motion of the wing and has therefore to
be estimated in the laboratory frame. Thus, at first or-
der, a reasonable corrected approximation for this term
is 4
3π ξnl(a + Aω)2.
|
1703.03926 | 1 | 1703 | 2017-03-11T07:48:34 | Osmotaxis in Escherichia coli through changes in motor speed | [
"physics.bio-ph"
] | Bacterial motility, and in particular repulsion or attraction towards specific chemicals, has been a subject of investigation for over 100 years, resulting in detailed understanding of bacterial chemotaxis and the corresponding sensory network in many bacterial species. For Escherichia coli most of the current understanding comes from the experiments with low levels of chemotactically-active ligands. However, chemotactically-inactive chemical species at concentrations found in the human gastrointestinal tract produce significant changes in E. coli's osmotic pressure, and have been shown to lead to taxis. To understand how these nonspecific physical signals influence motility, we look at the response of individual bacterial flagellar motors under step-wise changes in external osmolarity. We combine these measurements with a population swimming assay under the same conditions. Unlike for chemotactic response, a long-term increase in swimming/motor speeds is observed, and in the motor rotational bias, both of which scale with the osmotic shock magnitude. We discuss how the speed changes we observe can lead to steady state bacterial accumulation. | physics.bio-ph | physics |
Osmotaxis in Escherichia coli through changes in motor
speed
Jerko Rosko1, Vincent Martinez2, Wilson Poon2 and Teuta Pilizota1
March 14, 2017
1 Centre for Synthetic and Systems Biology, Institute of Cell Biology, School of Biological
Sciences, University of Edinburgh, Alexander Crum Brown Road, EH9 3FF, Edinburgh, UK
2 Scottish Universities Physics Alliance and School of Physics and Astronomy, The University
of Edinburgh, JCMB, Peter Guthrie Tait Road, Edinburgh EH9 3FD
Abstract
Bacterial motility, and in particular repulsion or attraction towards specific chemicals,
has been a subject of investigation for over 100 years, resulting in detailed understanding of
bacterial chemotaxis and the corresponding sensory network in many bacterial species. For
Escherichia coli most of the current understanding comes from the experiments with low levels
of chemotactically-active ligands. However, chemotactically-inactive chemical species at con-
centrations found in the human gastrointestinal tract produce significant changes in E. coli's
osmotic pressure, and have been shown to lead to taxis. To understand how these nonspecific
physical signals influence motility, we look at the response of individual bacterial flagellar
motors under step-wise changes in external osmolarity. We combine these measurements with
a population swimming assay under the same conditions. Unlike for chemotactic response, a
long term increase in swimming/motor speeds is observed, and in the motor rotational bias,
both of which scale with the osmotic shock magnitude. We discuss how the speed changes we
observe can lead to steady state bacterial accumulation.
Introduction
1
Many bacterial species are not only able to self propel (exhibit motility), but also direct their
motion towards more favorable environments. This behavior, called taxis [Krell et al., 2011, Pur-
cell, 1977], has long been a subject of scientific investigation, as it serves a variety of purposes:
seeking out nutrients and avoiding toxic substances [Wadhams and Armitage, 2004, Adler, 1969],
identifying thermal [Paster and Ryu, 2007] and oxygen [Adler et al., 2012] gradients, as well as aid-
ing pathogenic species in infecting their hosts [Rivera-Chávez et al., 2013, Cullender et al., 2013].
The understanding of bacterial taxis is not only important when it comes to bacterial motility
and accumulation; it also serves as a model for biological signal processing. Of particular interest
are the precision [Segal et al., 1986, Neumann et al., 2014], sensitivity [Cluzel et al., 2000] and
robustness [Yuan et al., 2012,Lele et al., 2012] that can be achieved with biological networks, and
potentially utilized for human design purposes [Navlakha and Bar-Joseph, 2014, Babaoglu et al.,
2006]. Specifically, bacterial chemotaxis, motion towards or away from specific chemicals [Wad-
hams and Armitage, 2004], was first described over 120 years ago [Massart, 1889]. Since then, the
systematic research efforts made it one of the best-studied topics in biology, especially when it
comes to Escherichia coli.
E. coli swims by rotating a bundle of flagellar filaments [Berg, 1973, Turner et al., 2000], each
powered by a bacterial flagellar motor (BFM), a rotary nano-machine that spins in the clockwise
(CW) or counter-clockwise (CCW) direction [Sowa and Berry, 2008]. Each bacterium possesses
several individual motors randomly distributed along the cell body [Tang and Blair, 1995, Turner
et al., 2000], which when rotating CCW enable formation of a stable filament bundle that propels
the cell forward [Berg, 2003]. When one or, most likely, a few motors switch to CW rotation, their
respective filaments fall out of the bundle, leading to a tumble event [Turner et al., 2000]. Forward
swimming, likely in a different direction due to Brownian rotation of the bacterial cell during the
tumble event, resumes when motors switch back to CCW direction and the bundle reforms [Berg,
1973, Turner et al., 2000].
1
The probability of switching increases with the intracellular concentration of phosphorylated
CheY protein (CheY-P) that interacts with the rotor of the BFM [Welch et al., 1993]. CheY-P
is a part of a feedback control circuit, the chemotactic network [Wadhams and Armitage, 2004],
which relays outside information to the motor and allows E. coli to direct its motion. Inputs of
the circuit are the methyl-accepting chemotactic proteins (MCPs), transmembrane proteins that
bind specific ligands in the cell exterior [Wadhams and Armitage, 2004], and through a signaling
cascade affect the CheY-P to CheY ratio. When sensing attractants or repellents in µM range, the
change in CheY-P to CheY ratio resets to the initial level within seconds, a characteristic feature of
the network termed perfect adaptation [Block et al., 1982]. Thus, directionality in the net motion
of the cell arises through transient tuning of motor switching frequency in response to external
stimuli [Wadhams and Armitage, 2004].
The majority of work on E. coli chemotaxis over the last 40 years has been performed in
a minimal phosphate buffer (termed Motility Buffer [Ryu et al., 2000]). However, one of the
primary habitats of E. coli is the gastrointestinal tract of humans and other warm-blooded animals
[Berg, 1996, Gordon and Cowling, 2003]. This complex environment features not only various
chemoattractants and repellents, but also spatial and temporal changes in osmolarity [Fordtran
and Locklear, 1966,Datta et al., 2016,Begley et al., 2005], which, in the stomach and small intestine
of humans, reach up to 400 mOsmol. The exact composition and osmolarity depend on the meal,
ingestion history and location within the gastrointestinal tract [Fordtran and Locklear, 1966].
Sudden osmotic increases, termed hyperosmotic shocks or upshocks, cause cell volume shrinkage
and require solute pumping and/or synthesis to re-inflate the cell and re-establish osmotic pressure
[Wood, 2015, Pilizota and Shaevitz, 2012]. Non-specific spatial taxis away from sources of high
concentrations, termed osmotaxis, has been observed in agar plates [Li et al., 1988]. Osmotic stimuli
can also send a signal down the network through mechanical stimulation of chemoreceptors [Vaknin
and Berg, 2006]; yet, osmotaxis was observed in gutted mutants lacking all chemotactic network
components [Li and Adler, 1993].
To clarify the exact nature of osmotactic response, we study the phenomenon on both single
cell and population levels, observing the rotation of individual flagellar motors under stepwise
increases in osmolarity, and measuring the swimming speeds of a population of ∼ 10000 bacteria
after exposure to an osmotic shock. The shock magnitudes we administer mimic those encountered
in the human gastrointestinal tract [Fordtran and Locklear, 1966]. We find that a stepwise increase
in osmolarity results in an elevated CCW-CW switching frequency that scales with the shock
magnitude. In addition, we observe osmokinesis post osmotic shock, i.e. significant changes in
the motor, and consequently, swimming speeds of bacteria. Lastly, for higher shock magnitudes,
we observe a loss of motor speed immediately post shock, followed by a transient, attractant-like
response that is coupled with a speed recovery. We discuss how the observed non-adaptive response
and osmokinesis can lead to taxis.
2 Results
2.1 Single motor response to an osmotic shock is complex
We begin with the reasonable assumption that at least one component of the chemotactic network
responds to an osmotic stimulus in order to generate previously observed osmotactic behavior [Li
et al., 1988]. Then, the response should be evident in the output of the chemotactic network, the
Clock-Wise (CW) bias of a single BFM. CW bias is defined as the fraction of time the BFM spends
rotating in the clockwise direction [Bai et al., 2010]:
CW Bias =
Ncw
Ntot
(1)
where Ncw is the number of data points corresponding to CW rotation and Ntot is the total number
of data points in a given time interval.
To compute the CW Bias, we measured the speed of an individual BFM when exposed to a
step-wise increase in the external osmolarity. SI Appendix Fig. 5 gives a schematic of the bead
assay used for measuring the motor speed [Ryu et al., 2000,Bai et al., 2010]. Briefly, we attach cells
to the cover slip and attach a latex bead, 0.5 µm in diameter, to a short filament stub [Ryu et al.,
2000, Bai et al., 2010]. The rotation of the BFM-driven bead is recorded using back-focal-plane
interferometry at a 10 kHz sampling rate [Pilizota et al., 2007, Denk and Webb, 1990, Svoboda
2
et al., 1993]. A representative single-motor rotation trace so obtained is shown in Fig. 1A. The
shaded light blue interval shows the motor speed and rotational direction prior to osmotic stimulus.
Positive motor speed values represent CCW rotation and negative CW. Fig. 1B shows the histogram
of the CW Bias obtained from the motor speed trace in A (Methods). Prior to the stimulus, the
motor switches from CCW to CW rotation ∼ 4 times per minute at CW Bias = 0.07, agreeing
with previous studies [Bai et al., 2010, Bai et al., 2013].
At t = 5min, the extracellular osmolarity is elevated by delivering 400 mM sucrose, at a local
flow rate of 0.68 µl/min [Buda et al., 2016], causing an osmotic upshock of 488 mOsmol/kg.
Immediately upon upshock, the motor stops switching rotational direction and speed drops. Then,
a recovery phase begins (Fig. 1A and B, non-shaded time interval) with the motor speed gradually
increasing while switching is absent. Post-recovery phase, which we define as the period after motor
switching has resumed, is characterized by an increased switching frequency. Consequentially, the
CW Bias is also elevated (to 0.16), and is maintained over at least 20 min, suggesting that, unlike
chemotaxis, the osmotactic response does not exhibit perfect adaptation [Segal et al., 1986, Block
et al., 1982]. The color map at the bottom of Fig. 1B is a compact representation of the histogram
in Fig. 1B.
To determine if CheY-P is necessary for the osmotic response observed in Fig. 1A and B,
we performed single-motor measurements on a strain lacking CheY (∆CheY mutant [Berg and
Turner, 1993, Fahren, 1995]). A representative trace is shown in Fig. 1C. The Recovery Phase of
the ∆CheY mutant is the same as observed in Fig. 1A for the chemotactic wild type. However, the
Post-recovery Phase of the mutant shows no switching events, indicating that CheY-P is necessary
for the elevated bias observed in Fig. 1A and B.
2.2 Osmotactic response does not exhibit perfect adaptation
We analyzed 69 cells exposed to three up-shock magnitudes to confirm the absence of perfect
adaptation observed in Post-recovery Phase (Fig. 1A and B). Up-shocks were delivered as in Fig. 1,
by exchanging VR Buffer (Methods) with the same buffer supplemented with 100, 200 and 400 mM
sucrose, corresponding to osmotic shocks of 111, 230 and 488 mOsmol/kg (similar to osmolalities
found in the small intestine [Fordtran and Locklear, 1966], see also SI Appendix for osmolality
measurements of all our buffers).
Fig. 2A shows color map histograms of each individual single-motor bias trace, where darker
color represents higher CW Bias and white represents smooth swimming. The Recovery Phase,
characterized by motor speed recovery and zero CW Bias period, scales with the shock magnitude.
SI Appendix Fig. 6 A shows the duration of the Recovery Phase, Trec,CW , against the shock
magnitude, where for the highest shock administered Trec,CW ∼5 min. The length of the Recovery
Phase roughly corresponds to the time E. coli takes to recover its volume and osmotic pressure
upon the hyperosmotic shock [Pilizota and Shaevitz, 2014]. Throughout the Post-Recovery Phase,
CW Bias levels do not, on average, relax to their initial pre-shock values. In our measurements this
phase lasts for ∼ 10-20 min, which is significantly longer than chemotactic adaptation times [Segal
et al., 1986] even when saturating attractant concentrations are used [Berg and Tedesco, 1975].
Evidence for this assertion is shown in the CW Bias histograms in Fig. 2B, which shows a
distribution computed from 5 min recordings in VR Buffer prior to shock, and Fig. 2C, which
shows a distribution computed from 3 min intervals taken at various time points, 12 or more
minutes after shock with 200 mM sucrose. Total of 120 single-motor recordings were used for the
pre-shock condition and 96 for after. The median value of the population CW Bias shifts from 0.01
pre-shock to 0.06 after addition of 200 mM sucrose. Fig. 2D shows the median CW Bias in time
for each of the three different shocks, calculated from the raw speeds of individual cells presented
in Fig. 2A, using a 60 s wide moving window. Here, only medians are shown for clarity and the
means, together with the interquartile range for each shock magnitude, are plotted in SI Appendix
Fig. 6 B-D. While the CW Bias shows some recovery in time, in particular for 111 mOsmol/kg
and 230 mOsmol/kg shocks, it proceeds on a slow time scale and the median values at the end of
our measurement time remain elevated with respect to the initial value.
A further corroboration of long-term increase in CW Bias post osmotic shock comes from
separating the CW Bias values of Fig. 2B according to the time point at which they were measured,
relative to the administration of the osmotic shock (t=0). This is displayed in the inset of Fig. 2B
and shows that the elevated CW Bias persists over a time scale as long as 1 h.
3
Figure 1: (A) An example 30 min speed trace obtained from a single BFM. The cell was initially
in VR Buffer (Methods), indicated in shaded light blue, and exposed to an osmotic upshock of
488 mOsmol/kg at t=5 min. The shock was delivered as a step increase by flowing in VRB
containing additional 400 mM Sucrose, resulting in shock magnitude of 488 mOsmol/kg. (B) A
histogram of clockwise bias computed by binning the trace in Fig. 1A into 30 second bins and
dividing the time spent rotating clockwise by the bin length (see also equation 1 and Methods).
Below is the same histogram condensed into a color map, with an intensity scale to the right. (C)
Single-motor speed trace of a ∆CheY mutant exposed to the same osmotic upshock as in Fig. 1A.
4
Figure 2: (A) Stacked single cell color maps (histograms) of CW Bias for three different shock
magnitudes (indicated on the left). Bin widths are 30 s. VR Buffer was exchanged for the same
buffer with the addition of sucrose at t=5 min. The white hatched column represents the period of
the media exchanged whose duration was ∼10-15 s. 22, 24 and 23 cells are given for the 111, 230
and 488 mOsmol/kg condition, respectively. Color map scale is given at the top. (B) Histogram
of CW Bias for cells prior to osmotic upshock (in VR Buffer) and (C) post osmotic upshock,
administered by exchanging VRB with VRB + 200 mM sucrose. Bin width is set to 0.020. Total
of 120 motors (each on a different cell) were used to construct Fig. 2B and 96 motors for Fig. 2C.
Inset in Fig. 2C plots these 96 single motor biases against the time after their respective osmotic
upshift. Gray shading represents the range between 25th and 75th percentile of the CW bias
distribution in VR Buffer given in B. (D) Median population CW Bias in time, computed from
cells given in Fig. 2A, for different shock magnitudes: 111 mOsmol/kg (red), 230 mOsmol/kg
(green) and 488 mOsmol/kg (dark blue). Black arrow indicates the time at which hyperosmotic
shock was administered.
5
00.570.27CW Bias111230Osmotic upshift [mOsmol/kg]488ACDB+111 mOsmol/kg+230 mOsmol/kg+488 mOsmol/kg2.3 Osmotic response shows osmokinesis, i.e. changes in motor speed
The response to an osmotic upshock includes not only CW Bias dynamics, but also changes in
motor and free-swimming speed. Fig. 3A shows a color map plot of normalized single-motor speeds
of the same 69 cells presented in Fig. 2A. All cells were originally in VR Buffer and subsequently
exposed to an osmotic upshock using 100, 200 or 400 mM sucrose, corresponding to upshifts of
111, 230 and 488 mOsmol/kg. BFM speeds were normalized with respect to the initial speed of
each motor, i.e. to the average value of the first 15 seconds of the recording.
Following the osmotic upshock, motors show two kinds of behavior. If a shock is of a large
magnitude, such as 488 mOsmol/kg, the speed drops sharply and significantly, with a phase of
speed recovery that follows. As can be seen from Fig. 3A and B the speed recovery (Trec,ω) lasts
∼4.4 min on average (5.2 min median). Additionally, this is of similar magnitude as Trec,CW
(5.4 min mean, 5.0 min median, SI Appendix Fig. 6) and likely corresponds to the period of
post-hyperosmotic shock volume recovery [Pilizota and Shaevitz, 2014]. After recovery, the speed
increase continues, leading to elevated levels compared to the pre-shock values, Fig. 3C. Weaker
upshocks, 111 and 230 mOsmol/kg, are characterized by an increase in motor speed without a
significant speed drop, as seen in Fig. 3A-C. The 0 mOsmol/kg condition is a buffer to buffer
control flush, where we used the same shocking protocol as for osmotic upshocks.
To explore the population-level significance of our observation of osmokinesis in single cells,
we next performed Differential Dynamic Microscopy (DDM) (see Methods) [Wilson et al., 2011,
Martinez et al., 2012]. DDM is a fast, high-throughput method for measuring the distribution of
swimming speeds in populations of a range of different self-propelled particles, averaging over up to
∼ 104 particles at the same time. The technique is well suited for rapid scanning of parameter space,
so that it also allowed us to extend the range of external osmolarities studied. DDM characterizes
the motility of a population of particles (in our case E. coli) by analyzing the statistics of temporal
fluctuations of pixel intensities in a sequence of low-optical resolution microscopy images, where the
intensity fluctuations are caused by the variation in number density of particles. Specifically, we
measure the differential image correlation function (DICF), which is effectively a power spectrum of
the difference between two images taken at separate time points [Wilson et al., 2011,Martinez et al.,
2012]. If a theoretical motility model exists, such as in the case of E. coli [Berg, 2003], the expected
DICF can be calculated and fitted to experimental data [Wilson et al., 2011], allowing accurate
estimates of the distribution of free-swimming speeds of a large number of bacteria [Martinez et al.,
2012].
Prior to DDM measurements cells were kept in VRB and osmotically shocked in a microfuge
tube. Upon the upshock cells were quickly placed into a capillary for DDM measurements and
the capillary was sealed, resulting in a fixed amount of oxygen present during the experiment.
Swimming speed recordings commenced within 2 min after the upshock (Methods) and are shown in
Fig. 3D. The gradual decrease of swimming speed with time observed in Fig. 3D for all magnitudes
of osmotic upshocks was previously characterized in Motility Buffer, where E. coli maintains Proton
Motive Force (PMF) using endogenous energy sources [Dawes and Ribbons, 1965, Schwarz-Linek
et al., 2016]. At fixed buffer composition, the time it takes to consume all available oxygen is
inversely proportional to the cell concentration, and upon oxygen exhaustion a sudden 'crash' in
swimming speed occurs [Schwarz-Linek et al., 2016]. Interestingly, we do not see such a 'crash'
in swimming speed for the lowest five values of the imposed upshock, but do see a 'crash' when
the upshock is at the highest value of 785 mOsmol/kg. Since the cell concentration is fixed, this
implies that cells consume oxygen at a significantly higher rate at the highest osmotic shock.
The increase in BFM speed observed in single cells, Fig. 3A, translates to an increase in popu-
lation swimming speed with increasing upshock magnitudes, Fig. 3D. Similarly, in agreement with
Fig. 3A, for high shock magnitudes, in particular for 473 mOsm/kg and 785 mOsm/kg sucrose
upshocks, a sharp speed drop is observed immediately upon upshock. We also see a Recovery
Phase, with duration increasing with the shock magnitude.
Increasing sucrose concentrations results in the increased viscosity of the media, consequently
increasing the drag coefficient on the motor (measured values of viscosity of our solutions are given
in SI Appendix). In Fig. 3 we show BFM speeds and cell swimming speeds uncorrected for this
effect, as the actual speed of the cell will be relevant for taxis and effective diffusivity. The viscosity
corrected BFM speeds, calculated under the assumption that the motor torque does not change
with the increasing viscosity, are shown in SI Appendix Fig. 8.
Furthermore, we check for the presence of steps during single motor speed recovery. To that
end, in SI Appendix Fig. 9 we show examples of BFM speeds during the Recovery Phase after
6
Figure 3: (A) Stacked single-motor (also single-cell) speed histograms, where the motor speed
for each BFM is normalized to the average value of the first 15 s. Bin widths are 15 s and the
color represents the bin height. Results are grouped by upshock magnitude, as indicated on the
left hand edge. The white hatched column represents the point where an osmotic shock was
performed by exchanging VR Buffer for VR Buffer + sucrose and the flow lasted for 10-15 s. 22,
24 and 23 cells are given for the 111, 230 and 488 mOsmol/kg conditions, respectively. The color
map scale is given at the top of the figure. (B) Time necessary to recover the average value of
pre-shock speed. Red horizontal bars are mean, and yellow triangles are median values, and the
graph contains 18, 22, 23 and 20 single motor data points for the 0, 111, 230 and 488 mOsmol/kg
upshocks. One value for the 230 mOsmol/kg condition and three for the 488 mOsmol/kg have
been excluded from the graph as these motors do not recover average initial speed in the course of
recording. The 0 mOsmol/kg condition is a buffer to buffer control flush. (C) Single motor speeds
calculated as 3 min averages corresponding to a section between t=17 and t=20 min in A. The
graph contains 12, 22, 24 and 23 single motor data points for the 0, 111, 230 and 488 mOsmol/kg
upshocks. The 0 mOsmol/kg condition contains 12 out of 18 control flushes that were at least
20 min long. (D) DDM measurement of swimming speeds following an osmotic shock. Cells were
shocked in microfuge tubes and brought into a microscope within 2 minutes. The legend shows
shock magnitudes and the mean speed is the average of swimming speeds obtained for each time
point in a range of different length scales (Methods). The systematic error of our measurements is
then calculated as the standard deviation of the mean, and falls within ∼ 5% of the mean value.
Here it was not plotted for clarity.
7
DNormalized motor speed051015202530Time [min]111230488 Osmotic upshift [mOsmol/kg]00.81.64365110215473785A2.4CBω17-20 [Hz]mOsmol/kgΔ[∙102mOsmol/kg]Δ[∙102mOsmol/kg]Trec,ω488 mOsmol/kg upshock, starting just after the osmotic shock was administered. Majority of the
traces do not show obvious steps during the BFM speed Recovery Phase.
3 Discussion
3.1 Origins of osmotaxis
With no external stimuli E.coli swims in an almost straight line and re-orients every so often in a
nearly-random fashion, performing a random walk with an effective diffusion constant D ∼ v2/α
[Schnitzer, 1993,Tailleur and Cates, 2008], where v is the swimming speed and α the tumbling rate
given by
3(1 − cos φ0)(τrun + τt)
α =
(2)
with τrun mean run time, φ0 (cid:118) 71◦ mean reorientation angle following a tumble, and τt the
τ 2
run
duration of the tumble event [Schnitzer, 1993, Lovely and Dahlquist, 1975].
Upon sensing a sudden increase in attractant concentration, E. coli sharply elevates the CCW
bias for ∼ 1 s and then slightly lowers it for ∼ 3 s, returning to the prestimulus level in ∼ 4 s [Block
et al., 1982]. It is this characteristic impulse response of modulating the tumbling rate that allows
E. coli to navigate towards a favorable environment. The detailed chemotactic performance and
its origins have often been studied [Schnitzer, 1993,Clark and Grant, 2005,de Gennes, 2004,Strong
et al., 1998, Cates, 2012, Schnitzer et al., 1990, Celani and Vergassola, 2010].
The observation that a chemotactic mutant of E. coli (CheRCheB) lacking normal methylation
and demetylation enzymes does not show chemotactic accumulation, but does respond to raised
attractant concentrations by lowering the tumble rate without adaptation [Block et al., 1982,Segal
et al., 1986] lead researchers to calculate the consequences of spatially varying α(x) and v(x) on
steady state bacterial density [Schnitzer, 1993, Schnitzer et al., 1990]. Similar calculations were
performed for particles with v and α dependent on the local density [Tailleur and Cates, 2008], or
when studying diffusion of bacteria in porous media [Licata et al., 2016]. These calculations offer
possible explanations for previously observed osmotaxis, given the characteristic speed and bias
response we here observed.
If we assume the simplest scenario, where there is no directional dependency of α(x) and v(x);
the steady state E. coli density will be inversely proportional to its swimming speed [Tailleur and
Cates, 2008]:
ρ(x) ∼ 1
v(x)
(3)
In Fig. 3C we show motor speeds, ω, and in Fig. 4 mean CCW motor interval (τ CCW ) and
tumbling rate (α) at a given osmolality. The ω and τ CCW were obtained from the last 3 min of
the post-osmotic shock recordings presented in Fig. 3A, which represent the long term osmolarity
dependent changes. We then calculated α from τ CCW and τ CW (Fig. 4 and SI Appendix Fig. 10)
using equation 2. The assumption was made that swimming speed v and the mean run time τ run
are proportional to the motor speed ω and the mean CCW interval τ CCW , in agreement with
previous studies [Magariyama et al., 1995], and our own DDM and motor measurements given in
Fig. 3.
Thus, based on experimental results we present here and previous theoretical calculations (equa-
tion 3 [Tailleur and Cates, 2008, Schnitzer, 1993]) we would expect accumulation at lower osmo-
larities, equivalent to negative taxis previously observed [Li et al., 1988]. The changes in tumbling
rates suggest differences in the time dependent approach to steady state, but not the steady state
density distribution itself [Schnitzer, 1993, Schnitzer et al., 1990].
The theoretical calculations we refer to make the assumption that spatial gradients of v(x)
and α(x) are small [Tailleur and Cates, 2008, Schnitzer, 1993], such that we do not expect local
gradients in steady state bacterial densities. For steeper osmotic gradients we expect a more
complex osmotactic response, in particular since the Recovery Phases shown in Fig. 2 and Fig. 3
will no longer be vanishingly short, as we assumed above, which could explain some previous
contradictory osmotactic observations [Li and Adler, 1993].
Our assumption that α(x) and v(x) lack directional dependence in the case of osmotaxis, could
be satisfied if the observed changes in α are not due to signaling within the chemotactic network, but
8
Figure 4: Mean counter clockwise motor interval, corresponding to cell runs, as a function of
osmotic upshift. The "0" condition box contains 96 mean run intervals, calculated by averaging
interval lengths over 5 min before osmotic upshift. Subsequent conditions, osmotic upshifts of 111,
230 and 488 mOsmol/kg, contain 21, 20 and 22 single-motor values calculated by averaging interval
lengths over the time span from t=12 min to t=15 min after an osmotic shock. Not all cells were
used to obtain τ CCW as note all cells had bound CCW intervals (see SI Appendix Fig. 11). In the
inset, we calculate the tumbling rate α according to equation 2, approximating the run times with
¯τccw and tumble times with ¯τcw given in SI Appendix Fig. 10.
9
purely due to changes in the interaction of CheY-P with the rotor units of the motor. For example,
previous observations indicate that at higher external osmolarities, osmotic pressure is kept the
same [Pilizota and Shaevitz, 2014]. Thus, the crowding in an already crowded cytoplasm [Parry
et al., 2014] increases, which can affect cytoplasmic interactions [Klumpp et al., 2013, Paudel and
Rueda, 2014] and the binding of CheY-P to the BFM, perhaps in a similar fashion to that observed
at higher hydrostatic pressures [Nishiyama et al., 2013].
However, it is also possible that the observed changes in τ CCW and ω are a consequence of
signal processing by the chemotactic network. In particular, Vaknin et al. reported that osmotically
induced changes in cell volume can perturb the chemoreceptors, and that the signal could travel
down some of the network components [Vaknin and Berg, 2006]. In such a scenario, α(x) and v(x)
would be directionally dependent (as in chemotactic response) and steady-state cell density would
need to be calculated taking into account the directional dependency as well as the characteristic
signal response we observed. Future work needs to investigate strains lacking specific parts of the
network to determine the contribution of signaling and different components of the network to the
osmotactic response.
3.2 Origins of osmokinesis
The large increase in motor speed post osmotic shock could be due to increases in PMF, possibly
through alterations in cell metabolism; or, as an alternative explanation, due to the increase in
number of stator units through mechanosensation [Lele et al., 2013,Tipping et al., 2013] or adaptive
motor remodeling [Yuan et al., 2012]. The BFM has been shown to act as a mechanosensor,
increasing the number of stator units in response to higher loads [Lele et al., 2013, Tipping et al.,
2013]. As the viscosity of the medium rises with addition of sucrose, the speed increase we observed
could be due to the incorporation of additional stator units. In fact, at higher loads, an increase in
the mean CW interval has been reported as well [Fahrner et al., 2003,Lele et al., 2013]. At our load,
a 0.5 µm bead attached to the motor via a short filament stub, the motor is still expected to operate
in the high-load, 'plateau' region of the torque-speed curve [Inoue et al., 2008, Lo, 2007] with the
estimated full stator number [Lo, 2007].
In addition, we observe an average 30 Hz increase in
motor-speeds even at our lowest osmotic shocks where the viscosity of the solution hardly changes
(it increases by 1.057 times) and even free-swimming cells with the motor operating in the linear-
torque regime do not increase the swimming speed at viscosity we use in our experiments [Martinez
et al., 2014]. Therefore, it is less likely that additional stator incorporation or adaptive motor
remodeling are sole explanations for the speed increases we observe.
SI Appendix Fig. 9 shows the motor speeds during the speed recovery phase for all of the
BFMs recorded in the +488 mOsmol/kg condition.
In majority of the traces no obvious steps
are observed, suggesting that the increase in motor speed could be due to the increase in PMF
with full set of stators present. Here we note that stator engagement with the rotor has been
reported as torque dependent [Tipping et al., 2013], where in absence of torque (motor rotation)
stators disengage from the rotor. We would then expect stator resurrection during the Recovery
Phase shown in SI Appendix Fig. 9. Absence of obvious step-wise increases indicates that volume
shrinkage caused by the osmotic shock could affect motor dynamics and perhaps prevent, or slow
down, stator disengagement.
The length of the motor Recovery Phase observed in Fig. 2 and 3 is ∼5-15 min, in line with
the volume recovery timescales observed previously post osmotic upshocks [Pilizota and Shaevitz,
2012, Pilizota and Shaevitz, 2014]. The timescales of speed recovery (at similar shock magni-
tudes) observed from DDM data in Fig. 3D are longer, with recovery lasting ∼40 min for the
473 mOsmol/kg upshock. Some variation in these times scales can be due to the difference be-
tween individual motor speed recovery and subsequent bundle formation. However, based on
Fig. 3D, we suspect that greater contribution to the difference comes from alterations in oxygen
consumption rate, which in turn shifts the time it takes to reach the maximum swimming speeds.
By looking at individual BFMs and population swimming speeds together, we reveal the main
characteristics of E. coli's motility response to step increases in external osmolarity. The response
consists of long term CW Bias and motor rotation/cell swimming speed increase. This is the
first observation of chemokinesis (osmokinesis) in E. coli. We discuss how such observed speed
increases can lead to negative taxis previously reported. Our study emphasizes the importance
of investigating bacterial motility in environments that mimic natural habitats, in the effort to
10
understand the role and evolutionary advantage swimming offers to bacterial cells [Lackraj et al.,
2016, Gauger et al., 2007, Tamar et al., 2016].
4 materials
4.1 E. coli strains and plasmids
E. coli strains KAF84 and KAF95 [Berg and Turner, 1993] were used for BFM speed and bias, and
MG1655 [Blattner et al., 1997] for DDM experiments. Both KAF84 and KAF 95 carry the fliC726
allele, (produce nonflagellate phenotypes), and contain a plasmid carrying an ampicillin resistance
and a fliC sticky gene (produces flagellar fillaments that stick readily to surfaces). Additionally,
KAF84 is a chemotactic wild type and KAF95 is a ∆CheY strain and therefore can not perform
chemotaxis, producing a smooth swimmer phenotype. MG1655 is a K-12 strain and a chemotactic
wild type.
4.2 E. coli Growth and culturing
KAF95, KAF84 and MG1655 cells were grown in Tryptone Broth (1% Bacto tryptone, 0.5% NaCl)
at 30◦C while shaken at 200 RPM [Bai et al., 2010, Martinez et al., 2012]. KAF95 and KAF84
were supplemented with 100 µg/ml of ampicillin and grown to OD=0.8-1.0 (Spectronic 200E Spec-
trophotometer, Thermo Scientific, USA), and MG1655 to OD=0.6. After growth cells were washed
into VR buffer (Volume Recovery) composed of the Modified Motility Buffer (MMB), which is
10 mM sodium phosphate buffer, pH=7.1 (an aqueous solution with 6.1mM of Na2HPO4, 3.9
mM of NaH2PO4) and 0.01 mM of Ethylenediaminetetraacetic acid (EDTA), with added Glycine
Betaine, Potassium Chloride and Choline Chloride to final concentrations of 10, 20 and 10 mM,
respectively. These compounds act as osmoprotectants and allow the cell to recover volume after
an osmotic shock. MMB is a variant of the Motility Buffer, commonly used in flagellar motor and
chemotaxis experiments [Ryu et al., 2000,Bai et al., 2010], with sodium phosphates substituted for
potassium phosphates. After washing, all the experiments were performed in VR buffer. KAF95
and KAF84 cells were washed by centrifuging them into a pellet and exchanging solution, while
MG1655 cells were washed by gentle filtration to preserve filaments [Schwarz-Linek et al., 2016]
and experiments were performed in VR buffer as for motor speed measurements.
4.3 Sample preparation and osmotic shock
For BFM experiments flagellar filaments were truncated by passing a bacterial suspension through
two syringes with narrow gauge needles (26G) connected with a plastic tube ('shearing device' [Bai
et al., 2010, Ryu et al., 2000, Pilizota et al., 2009]). Subsequently, cells with truncated filaments
were washed by centrifugation. Slides for BFM experiments were prepared as before [Bai et al.,
2010, Pilizota et al., 2009] by layering two parallel strips of double sided sticky tape onto a micro-
scope slide and covering them with a cover glass, forming a tunnel slide (SI Appendix Fig. 5) of
approximate volume of ∼ 8 µl. 1% poly-L-lysine was loaded into the tunnel and extensively washed
out after keeping it in for ∼10 s to allow glass coating. Cells with truncated filaments were loaded
into the tunnel and incubated for 10 min in a humid environment to prevent evaporation. Subse-
quently, non attached cells were washed out. Next, 0.5 µm beads in diameter (Polysciences) were
added and incubated for 10 min to allow sticking to the filaments and excess beads were washed out
post incubation. Osmotic shocks were performed while the slide was in the microscope by adding
24 µl of the shocking solution to one end and immediately bringing a piece of tissue paper to the
other, resulting in flow and exchange of media. The flow duration was approximately 10-15 s and
the local flow rate that the attached bacterium experienced was 0.68 µl/min [Buda et al., 2016].
Tunnel was sealed after the shock to prevent evaporation and thus potential further increase in
osmolarity throughout the course of the experiment. For DDM experiments an ∼400 µm deep
flat capillary (Vitrocom) glass sample was filled with ∼150 µl of bacterial suspensions immediately
after upshock, and subsequently sealed to prevent evaporation during the experiment.
4.4 Microscopy and data collection
For BFM experiments backfocal plane interferometry [Denk and Webb, 1990,Svoboda et al., 1993]
was performed using a custom built microscope and a 855 nm laser (Blue Sky Research, USA)
11
which formed a weak optical trap. The rotating bead attached to a flagellar stub was brought into
the focus of the laser and the back-focal plane of the condenser was imaged onto a position sensitive
detector (PSD Model 2931, New Focus, USA). The voltage signal from the PSD was passed through
an analog anti-aliasing filter (low pass, Bessel type filter with a cut off frequency of 2.5 kHz, Krohn-
Hite Corporation, USA) and sampled at 10 kHz (PCIE-6351 DAQ, National Instruments). For
DDM experiments imaging within the glass capillary was performed at 100 µm away from the
bottom of the capillary to avoid any interaction with the glass wall. The imaging begun within
∼ 2min post upshock, and consisted of a time-series of phase-contrast images (Nikon TE300 Eclipse
fitted with a Nikon Plan Fluor 10×Ph1 objective, NA=0.3, Ph1 phase-contrast illumination plate).
Imaging was performed at 100 Hz sampling rate (Mikrotron high-speed camera (MC 1362) and
frame grabber (Inspecta 5, 1-Gb memory)) for ∼ 40 s per movie duration and using 512x512 pixels
field of view. Both DDM and BFM experiments were performed at room temperature (21±1◦C).
4.5 Data analysis
Data collected during BFM experiments was analyzed in the following way. X and Y signals
(obtained from voltages from the PSD) were passed through a moving window discrete Fourier
transform, where the window moves point by point and the window size was 1.684 s. Time traces
of single motor speed obtained in such a way were processed to calculate the CW Bias using
Equation 1. CW Biases were calculated for a fixed interval sizes (30 s in Fig. 2A, 3 min in Fig. 2C
and 5 min in Fig.2B), and using a 60 s long moving window in Fig. 2D and SI Appendix Fig. 7B-E.
To calculate τ CCW we first take a mean of the τCCW intervals belonging to each individual BFM
rotation trace. We then pool so obtained mean cell intervals into a distribution and obtained
its mean (similar is true for τ CW , but in opposite direction). We include only bound CCW and
CW intervals (see Fig. 11). In DDM experiments, the mean swimming speed was calculated from
DDM movies as an average over ∼104 cells/ml. Details of the image processing and data analysis
were as before [Wilson et al., 2011,Martinez et al., 2012,Martinez et al., 2014,Schwarz-Linek et al.,
2016]. DDM allows the measurement of an advective speed, simultaneously, over a range of spatial-
frequency q, where each q defines a length-scale L=2*pi/q (for more details, see [Wilson et al.,
2011, Martinez et al., 2012]). A mean swimming speed is extracted by averaging over q. Data
plotted in Fig. 3D were obtained by averaging over the range 0.5 (cid:46) q (cid:46) 2.2 µm−1, corresponding
to a range in length-scale 3 (cid:46) L (cid:46) 13 µm.
5 Acknowledgments
This work was supported by Chancellor's Fellowship to TP. WCKP and VM were funded by an EP-
SRC Programme Grant EP/J007404/1 and an ERC advanced grant (AdG 340877 PHYSAPS). We
thank all of the members of the Pilizota lab, Poon Lab, and Alex Morozov and Filippo Menolascina
for useful discussions and support.
12
6 SUPPORTING INFORMATION
6.1 Growth Media Osmolarities
Growth media osmolalities were measured with an osmometer (Micro-Digital Osmometer MOD200
Plus, Camlab, Cambridge, UK) and are given in Table 1.
6.2 Motility Media Viscosities
Motility media viscosities were measured with a rheometer (TA Instruments AR2000, cone-plate
geometry, 60 cm, 0.5◦) and are given in Table 2.
6.3 Supporting Figures
Supporting Figures to the main text are given with the captions.
13
Table 1: Media osmolalities
Osmolality (mOsmol/kg)
Media
VRB
VRB+40 mM sucrose
VRB+60 mM sucrose
VRB+100 mM sucrose
VRB+200 mM sucrose
VRB+400 mM sucrose
VRB+600 mM sucrose
92
135
157
203
320
580
877
Table 2: Measured media viscosities (in agreement with listed in [Swindells, 1958])
Media
Motility Media
Motility Media + 300 mM sucrose
Motility Media + 600 mM sucrose
Motility Meida + 900 mM sucrose
Motility Media + 1850 mM sucrose
Viscosity (Pa s)
0.00112 ± 0.00001
0.00146 ± 0.00001
0.00192 ± 0.00002
0.00290 ± 0.00007
0.0172 ± 0.0001
Figure 5: Schematic representation of the bead assay and tunnel slide used for measurements. (A)
Genetically modified bacterial flagellar filament (FliCSticky [Fahren, 1995]) stub was attached to
a latex bead of 0.5 µm diameter. The bead was placed in a heavily attenuated optical trap and
the bead rotation recorded via back focal plane interferometry (Methods). (B) A tunnel slide,
formed by a double sided sticky tape separating a microscope slide and a glass cover slip, was used
to deliver osmotic shocks. The slide was sealed with vaseline upon the osmotic shock to prevent
evaporation.
14
Figure 6: (A) Recovery Phase duration (Trec,CW ), calculated as the time needed for the post-shock
median population bias to recover to the median VR Buffer bias (CW Bias=0.01, histogram in
main text Fig. 2B). Trec,CW was calculated for each of the three shock conditions, whose biases are
presented in panels C-D. The points represent times at which the post shock median reaches 0.01,
relative to the time of shock delivery. The error bars give the populational variability in terms of
recovery times for the 25th and the 75th percentile of population bias. (B) Buffer to buffer control
flush CW Bias during the course of the experiment is given. Biases were calculated on a moving
window of 60 s length and the VR Buffer was flown into the tunnel slide in the same way and
duration as for the osmotic shock experiments. 12 cells in total are presented, the solid black line
is the median, dashed is the mean and the shading represents the area between the 25th and 75th
percentile. (C) Same as B but for the 111 mOsmol/kg osmotic upshift and the plot represents the
median of the 22 single cell bias traces (thick red line), the mean (dashed black line) and the red
shading is the area between the 25th and the 75th percentile. (D) and (E) are the same as C, only
for the 230 mOsmol/kg (24 cells) and the 488 mOsmol/kg (23 cells) osmotic upshifts, respectively.
15
ATrec,cw[min]Δ[mOsmol/kg]BCDEFigure 7: Histogram of single-motor speeds for E. coli KAF84 in the Volume Recovery Buffer
(VRB). Each motor comes from a different single cell. Each speed is computed by taking the first
15 s of various recordings made in VRB, selecting the CCW speed from the trace and averaging
it. The histogram contains 118 cells and the mean value of the distribution is 116±2 Hz.
16
Figure 8: (A) Version of the main Fig. 3 where the motor and swimming speeds have been corrected
for sucrose viscosity, under the assumption that no additional stators are incorporated at this
increased load (see Discussion in the main text). Viscosity correction for 100, 200 and 400 mM
sucrose used are a multiplicative factor of 1.057, 1.195 and 1.444, respectively (measured and
listed in [Swindells, 1958]). Each line represents a single motor speed trace binned into 15 s
intervals where the color represents the height of the bin, normalized to the first bin. Results are
grouped by upshock magnitude, as indicated on the left hand edge. The white hatched column
represents the point where an osmotic shock was administered by exchanging VR Buffer for VR
Buffer + sucrose and the flow lasted for 10-15 s. There are 22, 24 and 23 cells for the 111, 230
and 488 mOsmol/kg condition, respectively. The color map scale is given at the top of the figure.
(B) DDM measurement of swimming speeds following an osmotic shock. Cells were shocked in
microfuge tubes and brought into a microscope within 2 min. The legend shows shock magnitudes
and the mean speed is the average of swimming speeds obtained for each time point in a range
of different length scales (Methods). The systematic error of our measurements is then calculated
as the standard deviation of the mean, and falls within ∼5% of the mean value (here not plotted
for clarity). The traces shown have been viscosity corrected as well, as free-swimming cells with
the motor operating in liner-torque regime do not increase the swimming speeds with viscosities
we used in our experiments [Martinez et al., 2014]. (C) Viscosity corrected single motors speeds
calculated as 3 min averages corresponding to a section between t=17 and t=20 min in A. The
0 mOsmol/kg condition is sampled from a set of buffer to buffer control flushes which were at least
20 minutes long (12 out of 18 control flushes). Red horizontal bars represent mean and yellow
triangles median values.
17
Normalized motor speed051015202530Time [min]111230488 Osmotic upshift [mOsmol/kg]0.41.22.0AC2.80Δ[∙102mOsmol/kg]ω17-20 [Hz]4365110215473785mOsmol/kgB050100150200Time [min]02468101214Swimming speed [μm/s]Figure 9: Traces of motor rotation after addition of 400 mM sucrose (all cells given in Fig. 2A and
Fig. 3A of the main text are given). We do not plot the initial motor speed, and start at the point
of speed drop due to an osmotic shock. Motor speeds were obtained as described in Methods.
18
Figure 10: Top: Mean CW interval distribution (τ CW ) for cells in Volume Recovery Buffer. Motor
rotation of 102 cells was sampled for 5 min and intervals counted as illustrated in SI Appendix
Fig. 11. Bottom: Mean CW interval distribution for motors where the sampling interval (3 min)
begins 12 or more minutes after an osmotic shock with 200 mM sucrose in VRB (+230 mOsmol/kg).
The inset contains data from the histogram plotted against the start time of the sampling interval
and t=0 corresponds to the osmotic shock.
19
Figure 11: A portion of a single-motor speed trace illustrating the counting of interval lengths
within a given sampling window. Positive speeds correspond to CCW rotation and negative to
CW rotation. Intervals i1 and i7 are not included when calculating the mean CCW interval due to
uncertainty in their beginning or end. Only intervals bounded by two CW intervals are counted.
Similar holds true when calculating mean CW interval, only those bounded by two CCW intervals
were included.
20
i1i7References
[Adler, 1969] Adler, J. (1969). Chemoreceptors in bacteria. Science, 166(3913):1588–1597.
[Adler et al., 2012] Adler, M., Erickstad, M., Gutierrez, E., and Groisman, A. (2012). Studies of
bacterial aerotaxis in a microfluidic device. Lab Chip, 12(22):4835–4847.
[Babaoglu et al., 2006] Babaoglu, O., Canright, G., Deutsch, A., Di Caro, G., Ducatelle, F., Gam-
bardella, L., Ganguly, N., Jelasity, M., Montemanni, R., Montresor, A., and Urnes, T. (2006).
Design patterns from biology for distributed computing. ACM Trans. Auton. Adapt. Syst.,
1:26–66.
[Bai et al., 2010] Bai, F., Branch, R., Nicolau, D. J., Pilizota, T., Steel, B., Maini, P., and Berry,
R. (2010). Conformational spread as a mechanism for cooperativity in the bacterial flagellar
switch. Science, 327(5966):685–689.
[Bai et al., 2013] Bai, F., Che, Y., Kami-ike, N., Ma, Q., Minamino, T., Sowa, Y., and Namba, K.
(2013). Populational heterogeneity vs. temporal fluctuation in escherichia coli flagellar motor
switching. Biophys J, 105(9):2123–2129.
[Begley et al., 2005] Begley, M., Gahan, C., and Hill, C. (2005). The interaction between bacteria
and bile. FEMS Microbiol Rev, 29(4):625–651.
[Berg, 1973] Berg, H. (1973). Bacteria swim by rotating their flagellar filaments. Nature, 245:380–
382.
[Berg, 2003] Berg, H. (2003). The rotary motor of bacterial flagella. Annual Review of Biochem-
istry, 72:19–54.
[Berg and Tedesco, 1975] Berg, H. C. and Tedesco, P. (1975). Transient response to chemotactic
stimuli in escherichia coli. Proc Natl Acad Sci USA, 72(8):3235–3239.
[Berg and Turner, 1993] Berg, H. C. and Turner, L. (1993). Torque generated by the flagellar
motor of escherichia coli. Biophys J, 65(5):2201–2216.
[Berg, 1996] Berg, R. (1996). The indigenous gastrointestinal microflora. Trends Microbiol,
4(11):430–435.
[Blattner et al., 1997] Blattner, F., Plunkett, G. r., Bloch, C., Perna, N., Burland, V., Riley,
M., Collado-Vides, J., Glasner, J., Rode, C., Mayhew, G., Gregor, J., Davis, N., Kirkpatrick,
H., Goeden, M., Rose, D., Mau, B., and Shao, Y. (1997). The complete genome sequence of
escherichia coli k-12. Science, 277(5331):1453–1462.
[Block et al., 1982] Block, S., Segall, J., and Berg, H. (1982).
chemotaxis. Cell, 31(1):215–226.
Impulse responses in bacterial
[Buda et al., 2016] Buda, R., Liu, Y., Yang, J., Hegde, S., Stevenson, K., Bai, F., and Pilizota,
T. (2016). Dynamics of escherichia coli's passive response to a sudden decreases in external
osmolarity. Proc Natl Acad Sci USA, September:doi:10.1073/pnas.1522185113.
[Cates, 2012] Cates, M. (2012). Diffusive transport without detailed balance in motile bacteria:
does microbiology need statistical physics? Rep Prog Phys, 75(4):042601.
[Celani and Vergassola, 2010] Celani, A. and Vergassola, M. (2010). Bacterial strategies for chemo-
taxis response. Proc Natl Acad Sci USA, 107(4):1391–1396.
[Clark and Grant, 2005] Clark, D. and Grant, L. (2005). The bacterial chemotactic response re-
flects a compromise between transient and steady-state behavior. Proc Natl Acad Sci USA,
102(26):9150–9155.
[Cluzel et al., 2000] Cluzel, P., Surette, M., and Leibler, S. (2000). An ultrasensitive bacterial
motor revealed by monitoring signaling proteins in single cells. Science, 287(5458):1652–1655.
21
[Cullender et al., 2013] Cullender, T., Chassaing, B., Janzon, A., Kumar, K., Muller, C., Werner,
J., Angenent, L., Bell, M., Hay, A., Peterson, D., Walter, J., Vijay-Kumar, M., Gewirtz, A., and
Ley, R. (2013). Innate and adaptive immunity interact to quench microbiome flagellar motility
in the gut. Cell Host Microbe, 14(15):571–581.
[Datta et al., 2016] Datta, S., Preska Steinberg, A., and Ismagilov, R. (2016). Polymers in the gut
compress the colonic mucus hydrogel. Proc Natl Acad Sci USA, 113(26):7041–7046.
[Dawes and Ribbons, 1965] Dawes, E. and Ribbons, D. (1965).
metabolism of escherichia coli. Biochem J., 95:332–343.
Studies on the endogenous
[de Gennes, 2004] de Gennes, P. (2004). Chemotaxis: the role of internal delays. Proc Natl Acad
Sci USA, 33(8):691–693.
[Denk and Webb, 1990] Denk, W. and Webb, W. (1990). Optical measurement of picometer dis-
placements of transparent microscopic objects. Appl Opt., 29(16):2382–2891.
[Fahren, 1995] Fahren, K. A. (1995). Studies of bacterial flagellar motors and filaments. Ph.D.
thesis, Harvard University, Cambridge, MA.
[Fahrner et al., 2003] Fahrner, K., Ryu, W., and Berg, H. (2003). Biomechanics: Bacterial flagellar
switching under load. Nature, 423:938–938.
[Fordtran and Locklear, 1966] Fordtran, J. and Locklear, T. (1966). Ionic constituents and osmo-
lality of gastric and small-intestinal fluids after eating. Am J Dig Dis, 11(7):503–521.
[Gauger et al., 2007] Gauger, E., Leatham, M., Mercado-Lubo, R., Laux, D., Conway, T., and
Cohen, P. (2007). Role of motility and the flhdc operon in escherichia coli mg1655 colonization
of the mouse intestine. Infect Immun, 75(7):3315–24.
[Gordon and Cowling, 2003] Gordon, D. and Cowling, A. (2003). The distribution and genetic
structure of escherichia coli in australian vertebrates: host and geographic effects. Microbiology,
149:3575–3586.
[Inoue et al., 2008] Inoue, Y., Lo, C., Fukuoka, H., Takahashi, H., Sowa, Y., Pilizota, T., Wad-
hams, G., Homma, M., Berry, R., and Ishijima, A. (2008). Torque–speed relationships of na+-
driven chimeric flagellar motors in escherichia coli. J. Mol. Biol, 376(5):1251–1259.
[Klumpp et al., 2013] Klumpp, S., Scott, M., Pedersen, S., and Hwa, T. (2013). Molecular crowd-
ing limits translation and cell growth. Proc Natl Acad Sci USA, 110(42):16754–16759.
[Krell et al., 2011] Krell, T., Lacal, J., Muñoz-Martínez, F., Reyes-Darias, J., Cadirci, B., García-
Fontana, C., and Ramos, J. (2011). Diversity at its best: bacterial taxis. Environmen. Microbiol.,
13(5):1115–1124.
[Lackraj et al., 2016] Lackraj, T., Kim, J., Tran, S., and Barnett Foster, D. (2016). Differential
modulation of flagella expression in enterohaemorrhagic escherichia coli o157: h7 by intestinal
short-chain fatty acid mixes. Microbiology, 162(10):1761–1772.
[Lele et al., 2012] Lele, P., Branch, R., Nathan, V., and Berg, H. (2012). Mechanism for adaptive
remodeling of the bacterial flagellar switch. Proc Natl Acad Sci USA, 109(49):20018–20022.
[Lele et al., 2013] Lele, P., Hosu, B., and Berg, H. (2013). Dynamics of mechanosensing in the
bacterial flagellar motor. Proc Natl Acad Sci USA, 110(29):11839–44.
[Li and Adler, 1993] Li, C. and Adler, J. (1993). Escherichia coli shows two types of behavioral
responses to osmotic upshift. J Bacteriol, 175(9):2564–25675.
[Li et al., 1988] Li, C., Boileau, A., Kung, C., and Adler, J. (1988). Osmotaxis in escherichia coli.
Proc. Natl. Acad. Sci. U.S.A., 85(24):9451–9455.
[Licata et al., 2016] Licata, N., Mohari, B., Fuqua, C., and Setayeshgar, S. (2016). Diffusion of
bacterial cells in porous media. Biophys. J, 110(1):247–257.
22
[Lo, 2007] Lo, C. (2007). Sodium energetics of chimeric flagellar motors in escherichia coli. Ph.D.
thesis. University of Oxford, Oxford, UK.
[Lovely and Dahlquist, 1975] Lovely, P. and Dahlquist, F. (1975). Statistical measures of bacterial
motility and chemotaxis. J Theor Biol, 50(2):477–496.
[Magariyama et al., 1995] Magariyama, Y., Sugiyama, S., Muramoto, K., Kawagishi, I., Imae,
Y., and Kudo, S. (1995). Simultaneous measurement of bacterial flagellar rotation rate and
swimming speed. Biophys. J, 69(5):2154–2162.
[Martinez et al., 2014] Martinez, V., Scwarz-Linek, J., Reufer, M., Wilson, L., Morozov, A., and
Poon, W. (2014). Flagellated bacterial motility in polymer solutions. Proc Natl Acad Sci USA,
111(50):17771–17776.
[Martinez et al., 2012] Martinez, V. A., Besseling, R., Croze, O., Tailleur, J., Reufer, M., Schwarz-
Linek, J., Wilson, L., Bees, M., and Poon, W. C. (2012). Differential dynamic microscopy: A
high-throughput method for characterizing the motility of microorganisms. Proc Natl Acad Sci
USA, 103(8):1637–1647.
[Massart, 1889] Massart, J. (1889). Sensibilité et adaptation des organismes à la concentration de
solutions salines. Arch. Biol., 9:515–570.
[Navlakha and Bar-Joseph, 2014] Navlakha, S. and Bar-Joseph, Z. (2014). Algorithms in nature:
the convergence of systems biology and computational thinking. Mol Syst Biol., 7:546–546.
[Neumann et al., 2014] Neumann, S., Vladimirov, N., Krembel, A., Wingreen, N., and Sourjik, V.
(2014). Adaptation in escherichia coli chemotaxis. PLoS ONE, 9(1):e84904.
[Nishiyama et al., 2013] Nishiyama, M., Sowa, Y., Kimura, Y., Homma, M., Ishijima, A., and
Terazima, M. (2013). High hydrostatic pressure induces counterclockwise to clockwise reversals
of the escherichia coli flagellar motor. J Bacteriol, 195(8):1809–1814.
[Parry et al., 2014] Parry, B., Surovtsev, I., Cabeen, M., O'Hern, C., Dufresne, E., and Jacobs-
Wagner, C. (2014). The bacterial cytoplasm has glass-like properties and is fluidized by metabolic
activity. Cell, 156(1-2):183–194.
[Paster and Ryu, 2007] Paster, E. and Ryu, W. (2007). The thermal impulse response of Es-
cherichia Coli. Proc Natl Acad Sci USA, 105(14):5373–5377.
[Paudel and Rueda, 2014] Paudel, B. and Rueda, D. (2014). Molecular crowding accelerates ri-
bozyme docking and catalysis. J Am Chem Soc, 136(48):16700–16703.
[Pilizota et al., 2007] Pilizota, T., Bilyard, T., Bai, F., Futai, M., Hosokawa, H., and Berry, R.
(2007). A programmable optical angle clamp for rotary molecular motors. Biophys J, 93(1):264–
275.
[Pilizota et al., 2009] Pilizota, T., Brown, M., Leake, M., Branch, R., Berry, R., and Armitage, J.
(2009). A molecular brake, not a clutch, stops the rhodobacter sphaeroides flagellar motor. Proc
Natl Acad Sci USA, 106(28):11582–11587.
[Pilizota and Shaevitz, 2012] Pilizota, T. and Shaevitz, J. (2012). Fast, multiphase volume adap-
tation to hyperosmotic shock by Escherichia coli. PLoS One, 7(4):e35205.
[Pilizota and Shaevitz, 2014] Pilizota, T. and Shaevitz, J. W. (2014). Origins of escherichia coli
growth rate and cell shape changes at high external osmolality. Biophys J, 107(8):1962–1969.
[Purcell, 1977] Purcell, E. (1977). Life at low reynolds number. American Journal of Physics,
45:3–11.
[Rivera-Chávez et al., 2013] Rivera-Chávez, F., Winter, S., Lopez, C., Xavier, M., Winter, M.,
Nuccio, S., Russell, J., Laughlin, R., Lawhon, S., Sterzenbach, T., Bevins, C., Tsolis, R., Harshey,
R., and Adams, LG ad Bäumler, A. (2013). Salmonella uses energy taxis to benefit from intestinal
inflammation. PLOS Pathog, 9(4):e1003267.
23
[Ryu et al., 2000] Ryu, W., Berry, R., and Berg, H. (2000). Torque-generating units of the flagellar
motor of escherichia coli have a high duty ratio. Nature, 403(6768):444–447.
[Schnitzer, 1993] Schnitzer, M. (1993). Theory of continuum random walks and application to
chemotaxis. Phys Rev E, 48(4):2553–2568.
[Schnitzer et al., 1990] Schnitzer, M., Block, S., Berg, H., and Purcell, E. (1990). Strategies for
chemotaxis. Symp Soc Gen Microbiol, 46:15–33.
[Schwarz-Linek et al., 2016] Schwarz-Linek, J., Arlt, J., Jepson, A., Dawson, A., Vissers, T.,
Miroli, D., Pilizota, T., Martinez, V., and Poon, W. (2016). Escherichia coli as a model ac-
tive colloid: A practical introduction. Colloids Surf B Biointerfaces, 137:2–16.
[Segal et al., 1986] Segal, J., Block, S., and Berg, H. (1986). Temporal comparisons in bacterial
chemotaxis. Proc Natl Acad Sci USA, 83(23):8987–8991.
[Sowa and Berry, 2008] Sowa, Y. and Berry, R. (2008). Bacterial flagellar motor. Quarterly Re-
views of Biophysics, 41(02):103–132.
[Strong et al., 1998] Strong, S., Freedman, B., Bialek, W., and Koberl, R. (1998). Adaptation and
optimal chemotactic strategy for e. coli. Phys Rev E, 57(4):4604–4617.
[Svoboda et al., 1993] Svoboda, K., Schmidt, C., Schnapp, B., and Block, S. (1993). Direct obser-
vation of kinesin stepping by optical trapping interferometry. Nature, 365(6448):721–727.
[Swindells, 1958] Swindells, J. (1958). Viscosities of sucrose solutions at various temperatures:
tables of recalculated values. United States, National Bureau of Standards.
[Tailleur and Cates, 2008] Tailleur, J. and Cates, M. (2008). Statistical mechanics of interacting
run-and-tumble bacteria. Phys Rev Lett, 100(21):218103.
[Tamar et al., 2016] Tamar, E., Koler, M., and Vaknin, A. (2016). The role of motility and chemo-
taxis in the bacterial colonization of protected surfaces. Sci Rep, 2016(6):19616.
[Tang and Blair, 1995] Tang, H. and Blair, D. (1995). Regulated underexpression of the flim
protein of escherichia coli and evidence for a location in the flagellar motor distinct from the
mota/motb torque generators. J Bacteriol, 177(12):3485–3495.
[Tipping et al., 2013] Tipping, M., Delalez, N., Lim, R., Berry, R., and Armitage, J. (2013). Load-
dependent assembly of the bacterial flagellar motor. mBio, 4:e00551–13.
[Turner et al., 2000] Turner, L., Ryu, W., and Berg, H. (2000). Real-time imaging of fluorescent
flagellar filaments. Journal of Bacteriology, 182(10):2793–2801.
[Vaknin and Berg, 2006] Vaknin, A. and Berg, H. (2006). Osmotic stress mechanically perturbs
chemoreceptors in escherichia coli. Proc Natl Acad Sci USA, 103(3):592–596.
[Wadhams and Armitage, 2004] Wadhams, G. and Armitage, J. (2004). Making sense of it all:
bacterial chemotaxis. Nat Rev Mol Cell Biol, 5(12):1024–1037.
[Welch et al., 1993] Welch, M., Oosawa, K., Aizawa, S.,
(1993).
Phosphorylation-dependent binding of a signal molecule to the flagellar switch of bacteria. Proc
Natl Acad Sci USA, 90(19):8787–8791.
and Eisenbach, M.
[Wilson et al., 2011] Wilson, L., Martinez, V., Schwarz-Linek, J., Tailleur, J., Bryant, G., Pusey,
P., and Poon, W. (2011). Differential dynamic microscopy of bacterial motility. Phys. Rev. Lett,
106(1):018101.
[Wood, 2015] Wood, J. (2015). Bacterial responses to osmotic challenges.
145(5):381–388.
J Gen Physiol,
[Yuan et al., 2012] Yuan, J., Branch, R., Hosu, B., and Berg, H. (2012). Adaptation at the output
of the chemotaxis signalling pathway. Nature, 484(7393):233–236.
24
|
1802.07169 | 1 | 1802 | 2018-02-20T16:07:17 | Self-organization principles of intracellular pattern formation | [
"physics.bio-ph",
"q-bio.SC"
] | Dynamic patterning of specific proteins is essential for the spatiotemporal regulation of many important intracellular processes in procaryotes, eucaryotes, and multicellular organisms. The emergence of patterns generated by interactions of diffusing proteins is a paradigmatic example for self-organization. In this article we review quantitative models for intracellular Min protein patterns in E. coli, Cdc42 polarization in S. cerevisiae, and the bipolar PAR protein patterns found in C. elegans. By analyzing the molecular processes driving these systems we derive a theoretical perspective on general principles underlying self-organized pattern formation. We argue that intracellular pattern formation is not captured by concepts such as "activators"', "inhibitors", or "substrate-depletion". Instead, intracellular pattern formation is based on the redistribution of proteins by cytosolic diffusion, and the cycling of proteins between distinct conformational states. Therefore, mass-conserving reaction-diffusion equations provide the most appropriate framework to study intracellular pattern formation. We conclude that directed transport, e.g. cytosolic diffusion along an actively maintained cytosolic gradient, is the key process underlying pattern formation. Thus the basic principle of self-organization is the establishment and maintenance of directed transport by intracellular protein dynamics. | physics.bio-ph | physics | Self-organization principles of intracellular pattern formation
J. Halatek. F. Brauns, and E. Frey∗
Arnold–Sommerfeld–Center for Theoretical Physics and Center for NanoScience,
Department of Physics, Ludwig-Maximilians-Universitat Munchen,
Theresienstrasse 37, D-80333 Munchen, Germany
(Dated: February 21, 2018)
Abstract
Dynamic patterning of specific proteins is essential for the spatiotemporal regulation of many
important intracellular processes in procaryotes, eucaryotes, and multicellular organisms. The
emergence of patterns generated by interactions of diffusing proteins is a paradigmatic example
for self-organization. In this article we review quantitative models for intracellular Min protein
patterns in E. coli, Cdc42 polarization in S. cerevisiae, and the bipolar PAR protein patterns
found in C. elegans. By analyzing the molecular processes driving these systems we derive a
theoretical perspective on general principles underlying self-organized pattern formation. We argue
that intracellular pattern formation is not captured by concepts such as "activators", "inhibitors",
or "substrate-depletion". Instead, intracellular pattern formation is based on the redistribution of
proteins by cytosolic diffusion, and the cycling of proteins between distinct conformational states.
Therefore, mass-conserving reaction-diffusion equations provide the most appropriate framework to
study intracellular pattern formation. We conclude that directed transport, e.g. cytosolic diffusion
along an actively maintained cytosolic gradient, is the key process underlying pattern formation.
Thus the basic principle of self-organization is the establishment and maintenance of directed
transport by intracellular protein dynamics.
Keywords:
self-organization; pattern formation;
intracellular patterns; reaction-diffusion; cell polarity;
NTPases
8
1
0
2
b
e
F
0
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
9
6
1
7
0
.
2
0
8
1
:
v
i
X
r
a
1
INTRODUCTION
In biological systems self-organization refers to the emergence of spatial and temporal
structure. Examples include the structure of the genetic code, the structure of proteins, the
structure of membrane and cytoplasm, or those of tissue, and connected neural networks.
On each of these levels, interactions resulting from the dynamics and structural comple-
mentarities of the system's constituents bring about the emergence of biological function.
Biological systems are the perfect example for the Aristotelian notion that "the whole is
more than the sum of it's parts". For centuries this phrase expressed nothing more than a
vague intuition that some set of organizational principles must underlie the complex phe-
nomena we observe around us. Owing to the advances in quantitative biology and theoretical
biological physics in recent decades, we have begun to understand how biological structure
and function originates from fundamental physical principles of self-organization. While we
are not yet in a position to define any universal physical principles of self-organization in
general, we are now able to identify recurring themes and principles in particular, important
areas like intracellular pattern formation. This will be the main focus of this review article.
The generic equilibrium state of any diffusion process is spatially uniform as diffusion re-
moves spatial gradients in chemical concentration. Self-organized pattern formation implies
that this equilibrium can be destabilized, such that an initially uniform system evolves to-
wards a non-uniform steady state - a pattern [1, 2]. Historically, the field of self-organized
pattern formation in chemical systems was initiated by Alan Turing in 1952 [1].
In his
seminal article on "The chemical basis of morphogenesis", Turing showed that the interplay
between molecular diffusion and chemical interactions can give rise to an instability of the
spatially uniform state. His general finding was that in a system with multiple reacting
components diffusing laterally on different time scales, the diffusive coupling itself can cause
an instability even if the system is in a stable chemical equilibrium. Turing was the first
to introduce a linear stability analysis for reaction-diffusion systems. To give the reader an
impression of the generality of his ideas let us briefly summarise the underlying mathemat-
ical concepts: The initial idea is that any random perturbation of a uniform steady state
can be decomposed in Fourier modes ∼ cos(qx) (Figure 1a). As long as amplitudes are
small, each of these modes grows or decays exponentially ∼ exp(σqt) cos(qx), depending on
the sign of the growth rate σq (or more precisely the real parts Re[σq]). By linear stability
2
analysis one computes the growth rates for modes with any wavenumber q. Turing found
that the interaction between chemical reactions and molecular diffusion can give rise to
bands of unstable modes with positive growth rates, i.e. situations where some modes with
particular wavelengths are amplified out of a random perturbation (Figure 1a). This gives
rise to pattern formation. In the following we will refer to the pattern forming instability as
lateral instability since it originates from lateral diffusive coupling.
As proof of principle Turing demonstrated this stability analysis for a general reaction-
diffusion system with two chemical components, but also discussed (oscillatory) cases with
three components. For 20 years his results received very little attention.
It was only in
1972 when Segel and Jackson [3] first interpreted Turing's linear stability analysis of the
two-component model, while Gierer and Meinhardt [4] in the same year proposed a num-
ber of specific two-component models and coined the terms "activator", "inhibitor", and
"activator–inhibitor mechanism" in this context.
Unfortunately, nowadays the terms "activator–inhibitor mechanism" and "Turing insta-
bility" are often thought to refer to identical concepts, despite the fact that the "activator–
inhibitor mechanism" only represents a particular interpretation of Turing's proof of princi-
ple analysis that is specific to some but not all two-component models. Furthermore, note
that Turing's general idea of a lateral instability is in fact not even limited to a particular
number of chemical components. In the literature the "activator–inhibitor mechanism" is
usually considered as a combination of "short range activation" and "long range inhibition"
or of "local activation" and "lateral inhibition" in order to convey the following heuristic
picture [2, 5, 6]:
Consider two chemical components. First – the (short range) activator – enhances its
own production in some autocatalytic fashion such that its concentration can increase expo-
nentially. If the diffusion coefficient of this component is small – any concentration peak will
only slowly disperse in the lateral direction. Secondly, the (long range) inhibitor, which is
also produced by the activator, but has a much large diffusion coefficient. Hence, it does not
accumulate locally with the activator but disperses laterally, where it inhibits the action of
the activator (see Fig. 1b). It is crucial to realize that this mechanisms is merely a heuristic
interpretation of a formal linear stability analysis presented by Turing. A quite common
misunderstanding in the biological literature is that pattern formation requires an activator
and an inhibitor. This does not in any way follow from the analysis by Turing [1], Segel and
3
Jackson [3], Gierer and Meinhardt [4–6], or any other analysis – activator–inhibitor models
are simply mathematically idealized examples of pattern forming systems. Moreover, the
underlying interpretation is actually restricted to systems with only two interacting chemical
components. Note that "chemical component" does not refer to a protein species, but to the
conformational state of a protein that determines its interactions with specific (conforma-
tions of) other proteins. Clearly, protein interaction networks include many conformational
states – not just two [7].
Furthermore, the activator–inhibitor interpretation inextricably links chemical properties
(e.g. autocatalytic action) to the diffusibility of the components (e.g. short range activation).
However, in the context of intracellular protein pattern formation the general distinction be-
tween diffusibilities is that between membrane-bound and cytosolic (conformational) states.
Accordingly, membrane-bound protein conformations would have to be considered as acti-
vators in the activator–inhibitor picture, and cytosolic protein conformation as inhibitors,
respectively. There are many reasons why this picture is not applicable to intracellular
protein dynamics – the most glaring discrepancy is that proteins are not produced auto-
catalytically on the membrane, which is the major (implicit) assumption underlying all
activator–inhibitor interpretations. As we will discuss in detail below, intracellular protein
pattern formation is generically independent of protein production and degradation (cf. [8]),
and intracellular protein dynamics are generically driven by the cycling of proteins between
membrane-bound and cytosolic conformations.
Another interpretation of Turing's mathematical analysis of two-component systems,
which appears to take these considerations into account, is the "activator–depletion" model
[4, 6] (see Fig. 1c). It differs from the activator–inhibitor model in making a specific choice
of the reaction terms and reinterpreting the rapidly diffusing component (formerly the in-
hibitor) as a substrate that is depleted by conversion into the activator. In this interpretation
the autocatalytic production (which increases activator and inhibitor concentrations) is re-
placed by an autocatalytic conversion of substrate to activator, which could be understood
as membrane attachment of a cytosolic protein. However, this type of model [4, 6] crucially
depends on cytosolic production of the substrate and degradation of the activator on the
membrane. In particular, concentration minima are not the result of a depleted cytosol (as
one might expect intuitively), but arise from the dominance of the activator degradation,
which effectively suppresses the autocatalytic conversion process [4, 6] (Fig. 1c). In other
4
words, accumulation of cytosolic proteins on the membrane is suppressed by concomitant
degradation of their membrane-bound forms.
Obviously, this assumption is highly specific and biologically implausible in terms of
intracellular protein dynamics. The use of a metalanguage with terms like "activator"
and "depletion" suggests that these concepts account for intracellular protein dynamics
where finite cytosolic particle pools play an important role. But the draining of finite
reservoirs is not actually the mechanism that drives pattern formation in activator–depletion
models. Like the activator–inhibitor model the activator–depletion model strictly depends
on production and degradation processes to explain pattern formation, and hence, it cannot
account for pattern formation in mass conserving systems. These issues clearly demonstrate
that heuristic interpretations and reinterpretations of specific mathematical models do not
generalize Turing's insight in a useful way. Indeed, "activator–inhibitor" and "activator–
depletion" models do not even provide a general picture of two-component systems, and
two-component systems are already a gross simplification of the biological reality. Hence,
there is no reason to assume that an intracellular pattern forming system must contain
activators and inhibitors, or involve depletion of substrates.
In our opinion, the use of
such metalanguage to describe the results of quantitative theoretical models does more
harm than good. It suggests that a unifying theoretical understanding is provided by the
idealized mathematical models to which intuitive terms like "activator", "inhibitor", and
"depletion" refer, whereas in reality, little is known about the actual general principles of
actual intracellular pattern formation.
Activator–inhibitor models do provide a legitimate phenomenological description of sys-
tems based on production (e.g. growth or gene regulation) and degradation, which is appli-
cable to some developmental phenomena [9] or vegetation patterns [10, 11]. However, even
in theses cases, it can be argued [10] that such models should be integrated in a more com-
plete modelling framework to account for specific details alone, rather than being treated as
paradigmatic models that convey the essence of pattern formation in general.
In this article, we provide a molecular perspective on intracellular pattern formation,
and review the underlying quantitative biological models, without reference to concepts
like "activators", "inhibitors", or depleting substrates. Instead, we review in the following
the specific implementation of pattern forming mechanisms by various protein interaction
systems – i.e.
the Min system in Escherischia coli, the Cdc42 system in Saccharomyces
5
cerevisiae, and the PAR system in Caenorhabditis elegans. Based on these systems we will
then extract and discuss recurring principles of the pattern forming dynamics, in particular
the fact that intracellular protein dynamics are based on cycling between different confor-
mational states. We conclude that intracellular pattern formation is, in essence, a spatial
redistribution process. Cytosolic concentration gradients are the primary means by which
directed transport is facilitated. The establishment and maintenance of such gradients is the
key principle underlying self-organized pattern formation. The proper theoretical framework
to study intracellular pattern formation is set by mass-conserving reaction-diffusion systems.
We will review recent theoretical advanced in this field [12] at the end of this article.
The Min system in E. coli
Cell division in E. coli requires a mechanism that reliably directs the assembly of the
Z-ring division machinery (FtsZ) to midcell [13]. How cells solve this task is one of the most
striking examples for intracellular pattern formation: the pole-to-pole Min protein oscillation
[14]. In the past two decades this system has been studied extensively both experimentally
[15–27] and theoretically [12, 27–31].
The Min protein system consists of three proteins, MinD, MinE, and MinC. In its ATP
bound form the ATPase MinD associates cooperatively with the cytoplasmic membrane
(see Fig. 2a). Membrane-bound MinD forms a complex with MinC, which inhibits Z-ring
assembly. Thus, to form a Z-ring at midcell, MinCD complexes must accumulate in the
polar zones of the cell but not at midcell. The dissociation of MinD from the membrane
is mediated by its ATPase Activating Protein (AAP) MinE, which is also recruited to the
membrane by MinD, forming MinDE complexes. MinE triggers the ATPase activity of
MinD initiating the detachment of both MinD-ADP and MinE. Subsequently, MinD-ADP
undergoes nucleotide exchange in the cytosol such that its ability to bind to the membrane
is restored (see Fig. 2a).
The joint action of MinD and MinE gives rise to oscillatory dynamics: MinD accumulates
at one cell pole, detaches due to the action of MinE, diffuses, and accumulates at the opposite
pole. The oscillation period is about 1 min, and during that time almost the entire mass of
MinD and MinE is redistributed through the cytosol from one end of the cell to the other
and back.
6
This example nicely illustrates the fact that pattern forming protein dynamics are in
essence protein redistribution processes [12, 30]. In other words, the emergent phenomenon
is directed transport, and not localized production and degradation (depletion), which serve
as the basis of activator–inhibitor (or activator–depletion) models.
It was suggested that binding of MinE to the membrane is essential for self-organized
pattern formation [23, 24, 32]. However, theses results were critically debated in the litera-
ture [33] and more recent experiments [34] have explicitly confirmed that MinE membrane
binding is not required for self-organized pattern formation. Therefore, we will not discuss
this process any further.
Furthermore, we note that Min protein oscillations are highly regular and therefore
amenable to a deterministic description.
Instances where stochastic effects [35] were re-
ported turned out to be overexpression artifacts [36] and not an indication for intrinsic noise
due to low copy numbers.
A striking lesson to be learned from the study of Min protein dynamics is the dependence
of the pattern forming process on cell geometry [14, 26, 27, 30, 37–39]. The pole-to-pole
oscillation in itself is a phenomenon intrinsically tied to the cell's geometry, which facilitates
the detection of a specific location in the cell. Over the past two decades a plethora of
fascinating observations has been made: (i) In filamentous cells, in which cell division is
inhibited, the pole-to-pole oscillation develops additional wave nodes showing that the Min
oscillation is a standing wave [14]. (ii) Experiments with nearly spherical cells show that, in
the majority of cases, the pattern forming process is able to detect the long axis, even though
it is much less pronounced than in wild-type, rod-shaped cells [37, 38]. (iii) In mutant cells
that were grown in nanofabricated chambers of various shapes a broad range of patterns has
been observed [26, 27]. In rectangular cells the oscillation can align with the long axis or
the short axis for the same dimensions of the cell. This shows that patterns with distinct
symmetries are stable under the same conditions: Min patterns are multistable.
The Cdc42 system in S. cerevisiae
Budding yeast (S. cerevisiae) cells are spherical and divide asymmetrically by growing a
daughter cell from a localized bud. The GTPase Cdc42 spatially coordinates bud formation
and growth via its downstream effectors. To that end, Cdc42 must accumulate within a
7
restricted region of the plasma membrane (a single Cdc42 cluster) [40]. Formation of a
Cdc42 cluster, i.e. cell polarization, is achieved in a self-organized fashion from a uniform
initial distribution even in the absence of spatial cues (symmetry breaking) [41].
Like all other GTPases, Cdc42 switches between an active GTP-bound state, and an
inactive GDP-bound state. Both active and inactive Cdc42 forms associate with the plasma
membrane, with Cdc42-GTP having the higher membrane affinity. Furthermore, Cdc42-
GDP is preferentially extracted from the membrane by its Guanine Nucleotide Dissociation
Inhibitor (GDI) Rdi1, which enables it to diffuse in the cytoplasm (see Fig. 2b) [42, 43].
Switching between GDP- and GTP-bound states is catalyzed by two classes of proteins:
Guanine nucleotide Exchange Factors (GEFs) catalyze the replacement of GDP by GTP,
switching Cdc42 to its active state; GTPase Activating Proteins (GAPs) enhance the slow
intrinsic GTPase activity of Cdc42, i.e. hydrolysis of GTP to GDP [44] (Note that owing
to their biochemical role, GAPs are called activating proteins, even though they switch
GTPases into their inactive, GDP-bound state. Moreover, these "activating proteins" are
in no way related to "activators" in the sense of "activator–inhibitor" models.). Cdc42 in
budding yeast has only one known GEF, Cdc24, and four GAPs: Bem2, Bem3, Rga1, and
Rga2. Further, a key player of the Cdc42 interaction network is the scaffold protein Bem1
which is recruited to the membrane by Cdc42-GTP, and itself recruits the GEF (Cdc24) to
form a Bem1–GEF complex (Fig. 2b) [45, 46].
Establishment and maintenance of Cdc42 polarization has been shown to rely on two
distinct and independent pathways of Cdc42 transport: (i) vesicle trafficking of vesicle-
bound Cdc42 along actin cables which require Cdc42-GTP (via its downstream effector
Bni1) [44, 47], and (ii) diffusive transport of GDI bound Cdc42 in the cytosol. Cytosolic
Cdc42 is recruited to the membrane by Bem1–GEF complexes (Fig. 2b) [45, 46, 48]. Either
of these transport pathways is sufficient for viability, as has been shown by either suppressing
vesicle trafficking (by depolymerizing actin) or inhibiting cytosolic diffusion of Cdc42-GDP
(by knocking out the GDI Rdi1) [47, 49, 50]. Since both pathways depend on Cdc42-
GTP, the pattern formation mechanism in both cases relies on polar activation (nucleotide
exchange) of Cdc42 by Bem1–GEF complexes, which are in turn recruited by Cdc42-GTP
[48, 51–54]. Various computational models of the Cdc42-Bem1-GEF interaction network
confirm that a positive feedback loop mediated by Bem1 is able to establish and maintain
polarization [55, 56].
8
Replacing Cdc42 with a constitutively active mutant suppresses GTPase cycling of Cdc42
and hence restricts it to membranes [47, 57]. Such mutants show that self-amplified directed
vesicle trafficking of Cdc42-GTP provides a viable self-organized polarization mechanism
in itself [57, 58]. Because the mutant Cdc42 is locked in its active state, these cells can
forego the polar activation of Cdc42 by Bem1–GEF complexes. Conceptual computational
models confirm that an actin mediated transport of Cdc42-GTP can in principle maintain
polarity [49, 57, 59–61], although studies of more realistic models show that key details of
the involved processes – endocytosis, exocytosis, and vesicle trafficking – are still unclear
[62, 63].
Interestingly, experiments where Bem1 was knocked out (or deprived of its ability to
recruit the GEF to active Cdc42) in cells with wild-type Cdc42 revealed that a third polar-
ization mechanism must exist, which is independent of both, Bem1 and vesicle trafficking
[64, 65]. Furthermore, polarization in the complete absence of Cdc42 transport has also been
observed [66], hinting at yet another pattern forming mechanism encoded within the inter-
action network of Cdc42. How these mechanisms operate independently of Bem1-mediated
feedback remains an open question that awaits experimental and theoretical analysis.
Normal cell division of budding yeast requires the reliable formation of a single bud-site,
i.e. a single Cdc42 cluster (polar zone, sometimes also called "polar cap"). Various mutant
strains exhibit initial transient formation of multiple Cdc42 clusters [67, 68], which then
compete for the limited total amount of Cdc42, leading to a "winner-takes-all" scenario
where only one cluster remains eventually [51, 69, 70].
The PAR system in C. elegans
So far we have discussed examples for intracellular pattern formation in unicellular
prokaryotes (Min) as well in eukaryotes (Cdc42). A well studied instance of intracellular
pattern formation in multicellular organisms is the establishment of the anterior-posterior
axis in the C. elegans zygote [71–74]. The key players here are two groups of PAR proteins:
The aPARs, PAR-3, PAR-6, and aPKC (atypical protein kinase C) localize in the anterior
half of the cell. The pPARs, PAR-1, PAR-2, and LGL, localize in the posterior half. In
the wild type, polarity is established upon fertilisation by cortical actomyosin flow oriented
towards the posterior centrosomes, in other words by active transport of pPAR proteins
9
[73, 75]. After polarity establishment this flow ceases, but polarity is maintained. In ad-
dition, it has been shown that polarity can be established without flow [75]. These results
suggest that PAR protein polarity in C. elegans is based on a reaction-diffusion mechanism.
The protein dynamics are based on the antagonism between membrane-bound aPAR and
pPAR proteins, mediated by mutual phosphorylation which initiates membrane detachment
at the interface between aPAR and pPAR domains near midcell (see Fig. 2c). Thus, PAR-
based pattern formation is driven by (mutual) detachment where opposing zones come into
contact, and is therefore quite different than the attachment (recruitment) based systems
discussed above.
Despite these apparent differences we will argue in the following sections that patterns
formation in all three systems is based on the same general principles.
GENERAL BIOPHYSICAL PRINCIPLES OF INTRACELLULAR PATTERN FOR-
MATION
Let us take a bird's eye view and ask: What are the general concepts and recurring
themes that are common to pattern formation in all of these biological systems?
In all cases the biological function associated with the respective pattern is mediated by
membrane-bound proteins alone, in other words: intracellular patterns are membrane-bound
patterns (exceptions are discussed further below). Furthermore, the diffusion coefficients of
membrane-bound proteins are generically at least two orders of magnitude lower than those
of their cytosolic counterparts, e.g. a typical value for diffusion along a membrane would
be between 0.01 µm2/s and 0.1 µm2/s, while a typical cytosolic protein has a diffusion
coefficient of about 10 µm2/s, e.g. see [66, 76].
The key unifying feature of all protein interaction systems is switching between different
protein states or conformations. The conformation (state) of a protein can change as a
consequence of interactions with other biomolecules (lipids, nucleotides, or other proteins).
Likewise, the interactions available to a protein are determined by its conformation. This can
be summarized as the switching paradigm of proteins (Fig. 2d), which is best exemplified
for NTPases such as MinD or Cdc42 whose dynamics are in essence driven by deactivation
and reactivation through nucleotide exchange. The phosphorylation and dephosphorylation
of PAR proteins by kinases and phosphatases, respectively, exemplifies the same principle.
10
In all these cases, switching is tied to membrane affinity, and thus to the flux of proteins
into and out of the cytosol.
Dynamics based on conformational switching conserve the copy number of the protein.
Therefore, intracellular protein dynamics are generically represented by mass-conserving
reaction-diffusion systems – and pattern formation in a mass-conserving system can only
be based on transport (redistribution), it cannot depend on production or degradation of
proteins. In the absence of active transport mechanisms (such as vesicle trafficking) the only
available transport process is molecular diffusion. Given that membrane-bound proteins
barely diffuse, we can assert that the biophysical role of the cytosol in these systems is that
of a (three-dimensional) 'transport layer'. Hence, the (functionally relevant) membrane-
bound protein pattern must originate from redistribution via the cytosol, i.e. the coupling
of membrane detachment in one spatial region of the cell to membrane attachment in another
region, through the maintenance of a diffusive flux in the cytosol.
However, transport by diffusion eliminates concentration gradients. Hence, if a diffusive
flux is to be maintained, a gradient needs to be sustained. Note that due to fast cytosolic
diffusion, this gradient can be rather shallow and still induce the flux necessary to establish
the pattern (the flux is simply given by the diffusion coefficient times the gradient).
Intracellular pattern formation is the localized accumulation of proteins on the membrane
by cytosolic redistribution. In this context, self-organization is the emergence of directed
transport that manifests itself as the formation of spatially separated attachment and de-
tachment zones, due to the interplay between cytosolic diffusion and protein interactions
(reactions).
Furthermore, we note that patterns can also be bound to other structures such as the
nucleoid [13] or the cytoskeleton [77–79]. In several of these pattern forming systems, the
dynamics of a nucleoid-bound, ParA-like ATPase results in different patterns, e.g. midcell
localization [80, 81] and pole-to-pole oscillations on the nucleoid of a single cargo as well
as equidistant positioning of multiple cargoes [82]. Various mechanisms to explain these
patterns have been proposed [83], including models that require ParA filament formation
[82, 84, 85] and ones which are based on a self-organizing concentration gradient of the
ATPase along the nucleoid [86–90]. In all these cases the dynamics of patterns are based
on the conformational switching of proteins between the cytosolic and the nucleoid-bound
(slowly diffusing) state. For midcell localization, the diffusive flux of the ATPase on the
11
nucleoid was found to be important [80, 91, 92].
Finally, we want to mention pattern forming systems that are based on a preexisting
template. The cell geometry itself is such a template, and macromolecules (phospholipids or
specific proteins) can have preferential affinity to accumulate in regions of specific membrane
curvature, e.g. the cell poles [93, 94]. However, theoretical analysis of Min protein dynamics
indicates that self-organized pattern formation based on lateral instability is robust against
heterogeneities in the membrane [95].
The elementary (i.e. most simple) intracellular pattern is cell polarization: the asym-
metric accumulation of proteins in a cell.
It lacks any intrinsic length scale and merely
serves to define a specific region of a cell (anterior/posterior domain, budding site). In the
following we will use cell polarity patterns as a paradigm to obtain a mechanistic picture of
intracellular pattern formation.
Cell polarity: The elementary pattern for intracellular pattern formation
How can one construct a general conceptual model for self-organized intracellular pat-
tern formation building on the principles discussed in the preceding section? The protein
dynamics are based on cycling of proteins between membrane-bound and cytosolic states,
and do not depend on production and degradation of proteins (cf. [8]). Hence, in a steady
state where the protein distribution is spatially uniform, attachment and detachment pro-
cesses must be balanced throughout the cell. Lateral instability, e.g. a Turing instability,
simply means that small spatial perturbations will be amplified. Let us therefore imagine
a perturbation of the density of a membrane-bound species where at some membrane po-
sition the protein concentration will be slightly larger than elsewhere on the membrane. If
this perturbation is to be amplified, proteins must be transported to the position where the
membrane density is (already) highest. For this transport of proteins, we only have cytosolic
diffusion at our disposal. To facilitate a directed transport to a specific position by cytosolic
diffusion, the cytosolic density at this position must itself be at a minimum (Fig. 3). In
order to reduce the cytosolic density at the position where the membrane density is highest,
the balance between attachment and detachment must shift in favour of attachment such
that protein mass flows from the cytosol to the membrane. Conversely, in the region where
the membrane density is lowest, the attachment-detachment balance must shift in favour of
12
detachment, thereby increasing the cytosolic density in this region. Only if changes in den-
sity shift attachment–detachment balance in this fashion will the initial perturbation on the
membrane be amplified by further attachment and detachment due to cytosolic transport
(Fig. 3).
To concisely summarize: Intracellular pattern formation can be understood as the forma-
tion of attachment and detachment zones, which are coupled through cytosolic gradients that
facilitate protein mass redistribution.
In this light, we will now look again at the specific biological systems introduced earlier,
and attempt to uncover the basic molecular mechanisms that lead to the formation and
maintenance of attachment and detachment zones.
QUANTITATIVE MODELS FOR INTRACELLULAR PATTERN FORMING SYS-
TEMS
The Cdc42 system in S. cerevisiae
The key interaction that drives self-organized Cdc42 polarization is the recruitment of
GEF by Cdc42, mediated by Bem1, giving rise to mutual recruitment of Cdc42, Bem1
and GEF. In a minimal model, the role of Bem1 and GEF can be summarized by an
effective Bem1–GEF complex, which is recruited to Cdc42-GTP on the membrane (Fig. 4a).
There, the GEF then recruits more Cdc42 from the cytosol and converts it into the GTP-
bound form. As a further simplification, recruitment and nucleotide exchange of Cdc42 by
Bem1–GEF complexes can be subsumed into a single step. Similarly hydrolysis of GTP
(catalyzed by GAPs) and extraction of CDC42 from the membrane can be conflated to a
single membrane dissociation step. Effectively, only active Cdc42 on the membrane and
inactive Cdc42 in the cytosol are considered (Fig. 4a).
How can this interaction scheme of mutual recruitment establish spatially separated at-
tachment and detachment zones for the polarity marker Cdc42? Cdc42-GTP on the mem-
brane acts as a "recruitment template" for Bem1–GEF complexes: a zone of high Cdc42-
GTP density on the membrane creates an attachment zone for Bem1, which in turn creates
an attachment zone for the GEF Cdc24, such that effectively Cdc42-GTP creates an at-
tachment zone for Bem1–GEF complexes (cf. panels (1) and (2) in Fig. 4b). A region of
13
high Bem1–GEF density on the membrane, in turn, acts as recruitment template for Cdc42,
creating a Cdc42 attachment zone, and locally enhances Cdc42 nucleotide exchange leading
to increased local Cdc42-GTP density (cf. panels (1') and (2') in Fig. 4b). In the absence of
Cdc42-GTP, very little Bem1 attaches to the membrane, such that detachment dominates.
Similarly, in the absence of GEF, Cdc42 is dominantly inactive, such that membrane extrac-
tion by GDI, i.e. detachment of Cdc42, dominates. Starting from a uniform state, a spatial
perturbation of either density will establish a mutual recruitment zone, with Cdc42–GTP
sustaining the attachment zone for Bem1–GEF, which in turn maintains the recruitment
and activation zone of Cdc42 (Fig. 4c).
Conceptually, cell polarization need not require two protein species that mutually re-
cruit each other: conceptual theoretical models for cell-polarity involving only two chemi-
cal components (effectively describing a single protein type in two conformational states –
membrane-bound and cytosolic) have also been studied [61, 96, 97]. In these models, patters
with multiple density peaks show "winner-takes-all" coarsening dynamics due to competi-
tion for the conserved total mass of proteins [96] (cf. the discussion of wavelength selection
and mass-conserving reaction–diffusion systems below).
The Min system in E. coli
Pole-to-pole oscillations are the result of interactions between MinD and MinE. In Fig. 5,
the key phases of the oscillation cycle are shown. Membrane-bound MinD facilitates further
accumulation of MinD and MinE on the membrane through recruitment. The recruitment
of MinD shifts the attachment-detachment balance towards further attachment. In contrast,
the recruitment of MinE shift this balance towards detachment. The structure of a polar
zone is such that MinE is accumulated in the form of MinDE complexes at its rim (sometimes
called the "E-ring"), whereas MinD accumulates at its tip (see Fig. 5 panels (2) and (4)).
A theoretical analysis of Min protein dynamics revealed that self-organized Min protein
pattern formation is based on two requirements [30]: The total copy number of MinD must
exceed that of MinE, and the recruitment rate of MinE must be larger than that of MinD.
The first condition ensures that all of the MinE can be bound as MinDE complexes on
the membrane (MinE-MinD detachment zone, panels (2) and (4) in Fig. 5) while leaving
a fraction of MinD free to initiate and maintain a MinD attachment zone. The second
14
condition causes MinE to be trapped immediately upon entry into a polar zone (MinD-MinE
attachment zone) and thereby "sequestrated" at the rim, creating a localized MinE-MinD
detachment zone (panels (1) and (3) in Fig. 5). Rebinding of MinE to the MinD-MinE
attachment zone at the tip of the polar zone is favoured due to the faster recruitment of
MinE and the fact that MinD is temporarily inactive after detachment. This leads to the
progressive conversion of attachment zones into detachment zones due to the shifting balance
in favor of MinDE complexes (detachment).
Below we will discuss that the inactivation of MinD upon MinE-stimulated hydrolysis
is essential for the establishment of intrinsic length scales and for the dependence of the
pattern–forming process on cell geometry (see also Fig. 7).
Biochemically the Min system of E. coli and the Cdc42 system of S. cerevisiea, are
closely related: both MinD and Cdc42 are NTPases regulated by enzymatic proteins such as
NTPase-activating proteins (NAPs), even though the regulation of Cdc42 activity is much
more complex. As we have argued above, the pattern forming dynamics of both systems
follow the same underlying physical principle: self-organized spatial separation of attachment
and detachment zones. Indeed, theoretical models have predicted that Min protein dynamics
can also give rise to stationary polar patterns [30]. Conversely, oscillations of Cdc42-marked
polar zones in budding yeast have been observed experimentally [98], while the non-spherical
fission yeast exhibits pole-to-pole oscillations of Cdc42 clusters during the polar growth phase
[99]. This raises the question whether there is a common underlying mechanism that unifies
Min protein patterns and Cdc42 polarization at the physical level.
The PAR system in C. elegans
PAR protein polarization in C. elegans is based on an antagonism between membrane-
bound aPAR and pPAR protein through mutual phosphorylation, cf. [75]. The major differ-
ence relative to the above discussed systems is the lack of an evident biochemical mechanism
for the formation of attachment zones, such as recruitment. Instead for C. elegans, attach-
ment zones result from mutual exclusion. In a zone with high aPAR concentration on the
membrane only aPAR can attach, as pPARs are immediately phosphorylated. Similarly, in
zones with high pPAR membrane concentration only pPAR can attach. At the interface be-
tween aPAR and pPAR zones each protein class drives the other off the membrane. Hence,
15
the interface acts as detachment zone for both aPAR and pPAR, whereas the anterior pole
(aPAR dominant) acts as aPAR attachment zone, and the posterior pole (pPAR) acts as
pPAR attachment zone. The key to pattern formation is the detachment zone, i.e. the
maintenance of the interface. There, aPAR and pPAR domains are actively separated from
each other, and cycling between the interface and the respective attachment zones maintains
the bi-polar pattern. Theoretical analysis [75] shows that a stable interface requires the rates
of the antagonistic interactions to be comparable.
A very interesting aspect of PAR protein pattern formation is the role of the cortical flow
[73, 75]. In the wild type it is used to segregate aPAR zones from pPAR zones, i.e. to form
the respective attachment zones. The polarized state is maintained after the flow ceases,
showing that maintenance of the interface is independent of the flow, i.e. it is self-organized.
ADVANCED INTRACELLULAR PATTERN FORMATION: WAVELENGTH SE-
LECTION, DEPENDENCE ON CELL GEOMETRY, AND MULTISTABILITY OF
PATTERNS
Pattern formation and length scale selection: the classical picture
Traditionally the phenomenon of self-organized pattern formation in reaction-diffusion
systems has been intrinsically linked to the postulated existence of a characteristic length
scale [2]. In particular, most authors define a Turing pattern as a pattern with a character-
istic length scale [100]. In our discussion so far such a length scale has only been mentioned
in passing as a phenomenon observed in specific E. coli mutants, and in budding yeast mu-
tants as a transient pattern of multiple Cdc42 clusters. Indeed, the theoretical analysis of all
quantitative models [30, 55, 71] discussed so far reveals that the existence of such a length
scale is in no way generic – despite the fact that all patterns emerge from a lateral instabil-
ity induced by diffusive coupling, i.e. a Turing instability. On the contrary, it appears that
the phenomenon biologists refer to as "the winner takes all" and physicist as "coarsening"
is the generic case, cf.
[27, 30, 101, 102]. Hence, the generic pattern is a polarized state
with a single concentration maximum and a single concentration minimum for each chemical
component irrespective of the system size.
In fact, the question of length scale selection is a highly nontrivial problem, and can only
16
be addressed in general by numerical simulations. Linear stability analysis (as introduced by
Turing) only predicts the length scale of the (transient) pattern that initially forms from the
uniform state (see Fig. 1a). This should not be confused with the length scale of the final
pattern. For instance, coarsening dynamics ("the winner takes all") are generic examples
for dynamics where the length scale collapses to the system size (or an intermediate length
scale) regardless of the initially selected length scale [103]. A case where the length scale
is predicted correctly by the linear stability analysis is when the growth of the pattern
saturates at small amplitude [2]. Some authors include saturation at small amplitude in
their definition of a Turing pattern [104]. While this definition is mathematically rigorous,
it is also very restrictive: For technical mathematical reasons this (supercritical bifurcation,
near threshold) case implies that the pattern must vanish if some system parameters are
slightly changed. If the pattern does not saturate at a very small amplitude early on, no
(reliable) prediction about the final pattern can be made based on the linear stability analysis
(except that some non-uniform pattern exists).
From the mathematical point of view this specific case (supercritical bifurcation, near
threshold) is very attractive as it lends itself to analytical calculations [2]. To readers less
interested in the mathematical details, these points may seem overly technical. However,
it is crucial to realize that – from the biological perspective – such technical limitations
(a pattern with very small amplitude (weak signal) that is highly sensitive to parameter
changes) imply patterns that are highly fragile (cf. [105]), and will therefore be eliminated
by natural selection. In that light it is not surprising that (robust) quantitative models of
biological systems, like those presented above, do not meet the constraints of small amplitude
and vicinity to a supercritical bifurcation. Partly because most mathematically motivated
work on pattern selection is based on the assumption that these constraints are met, very
little is known about pattern selection in (evolutionarily robust) biological systems.
Yet, several key aspects of pattern selection can be inferred from the theoretical analysis
of models for actual biological systems. For example, theoretical analysis of Min protein
patterns [30] showed that standing wave patterns with a finite wavelength emerge if the lat-
eral redistribution of Min proteins is canalized, see Fig. 7. In terms of the general principles
discussed above, this means that attachment and detachment zones are strongly coupled
through cytosolic transport. The flux off the membrane in a detachment zone is of the same
order of magnitude as the flux onto the membrane in the attachment zone. Hence, the frac-
17
tion of cytosolic proteins remains approximately constant during the redistribution process
(Fig. 7a). As we have discussed, attachment and detachment zones are regulated by the
membrane kinetics of the specific biochemical model, while transport depends on cytosolic
kinetics and diffusion. Canalized transfer leads to the emergence of a characteristic sepa-
ration distance between attachment and detachment zones which depends in a nontrivial
manner on the system parameters (Fig. 7a). The particular parameter dependence of such
characteristic redistribution length scales remains an open question for the Min system, and
even more so for general reaction diffusion systems. However, it was demonstrated that the
total mass flux due to canalized transfer can be inferred from the linear stability analysis
for the Min model [30]. The flux coupling (cytosolic exchange) between detachment and at-
tachment zones is weak, if a cytosolic reservoir is filled and depleted during detachment- and
attachment-dominant phases, respectively (see Fig. 7b). It seems intuitive that a redistribu-
tion process through a "well-mixed" cytosolic reservoir does not dictate an intrinsic length
scale for pattern forming dynamics. Moreover, the analysis of quantitative models (such as
the Min model) does provide strong evidence that length scale selection in reaction-diffusion
systems essentially relates to the length scales of directed ("canalized") transport. However,
the precise details underlying the emergence of intrinsic length scales remain unknown.
Cell geometry and pattern selection
In the previous section we discussed why it is important to consider the seemingly tech-
nical limitations underlying some mathematical results about pattern formation in order
to correctly understand quantitative biological models for intracellular pattern formation.
Besides the question of length scale selection, the effect of system geometry (i.e. cell shape)
is in this respect another case in point.
At first, linear stability analysis of reaction-diffusion systems was exclusively restricted
to planar geometries such as lines and flat surfaces. Only very recently, the method was
extended to account for (2d) circular geometries where dynamics can take place on the
boundary of the circle (membrane) as well as in the bulk (cytosol), and proteins exchange
between the two domains (membrane-cytosol cycling) [106]. This important first step to-
wards quantitative modeling of intracellular protein dynamics was still limited to purely
linear attachment-detachment dynamics (thus excluding the cases of cooperative attach-
18
ment, recruitment, or antagonistic detachment). Later, linear stability analysis methods
were extended to general attachment-detachment kinetics by for (3d) spherical geometry
[55], and for (2d) elliptical geometry [30].
The extension to elliptical geometry revealed a very important point [30]:
it is in no
way generic that patterns align with the long axis of a cell, i.e.
there is in general no
intrinsic preference for the selection of long axis patterns over short axis patterns. From a
biological perspective this is a crucial finding, since proper axis selection is typically linked
to the spatial nature of the biological function mediated by the pattern in the first place.
Intuitively, one may expect that axis selection is connected to the "characteristic length
scale" of the pattern obtained from a linear stability analysis in a planar (flat) geometry:
whichever axis length of the cell is closer to this "characteristic length scale" determines the
axis that is selected. The intuition behind this is that the pattern has "to fit" into the cell.
So far, however, there is no evidence to support this intuition. On the contrary, it appears
that the question of axis selection is much more complicated. In a study combining theory
and experiments Wu et al.
[26, 27] analyzed the Min protein patterns in rectangular cell
geometries of various sizes and aspect ratios. The experiments found that a broad range
of patterns (aligned with the long axis or the short axis) can emerge in the same system
geometries. Hence, intracellular Min protein patterns are multistable, and can conform to
a variety of intrinsic length scales instead of one "characteristic" length scale.
Theoretical analysis [27] confirmed multistability of Min patterns, and was able to link all
observations to the emergence of an intrinsic length scale for diffusive cytosolic redistribution
("canalized transport") in the model: the stronger the flux-coupling between attachment
and detachment zones, the stronger was the dependence of the pattern-forming process on
cell geometry, and the greater the number of multistable patterns with distinct symmetries
(long axis or short axis alignment) observed in a broad range of rectangular cell geometries
[27]. This strongly suggest that pattern selection and the influence of cell geometry are –
just like wavelength selection – closely tied to the length scale of lateral transport and the
strength of the coupling between attachment and detachment zones. Note that this is a
very different picture from the one presented by "activator–inhibitor" models. In the latter,
the length scale is set by the degradation and production length scale, e.g. the length scale
over which autocatalytic production of the slowly diffusing component (activator) dominates
over its own degradation. In contrast, for intracellular protein dynamics the essential length
19
scale appears to be the (mean) distance over which membrane-bound proteins (the slowly
diffusing components) are redistributed in the (fast diffusing) cytosolic state following their
detachment.
We will next discuss how the switch-like behaviour of proteins appears to be essential for
the emergence and regulation of such transport length scales.
The different roles of cytosolic kinetics and membrane kinetics
As we discussed above, the switch-like behaviour of proteins between active and inactive
states is a central paradigm of protein dynamics.
In many cases the switch alters the
affinity of proteins for the membrane and can thus be utilized to stimulate attachment or
detachment.
In case of the Min system in E. coli only active MinD-ATP can bind and
be recruited to the membrane. As theoretical analysis [30] has shown, this property is
essential for the regulation of intracellular transport and the establishment of "canalized
transfer": Since MinD detaches from the membrane in the inactive MinD-ADP form it
cannot immediately rebind. Hence, if the timescale of cytosolic reactivation (nucleotide
exchange) is sufficiently long, a MinD protein detaching from a polar zone with high MinD
membrane density can leave the polar zone, by diffusion, before being reactivated. In this
way, rebinding of detached MinD to polar zones can be suppressed even if the affinity of
cytosolic MinD-ATP for membrane-bound MinD is very high. In fact, a very high affinity for
membrane-bound MinD can serve to promote rebinding of MinD-ATP in new (weak) polar
zones. In other words: to promote directed transport of MinD from high (MinD and MinE
membrane) density region to a low density region.
In this way, the high density region
becomes a detachment zone and the low density region an attachment zone.
Increasing
the affinity of cytosolic MinD-ATP to membrane-bound MinD (recruitment rate) simply
increases the attachment in the low density region – but not in the high density region
where MinD detached and cannot rebind due to delayed nucleotide exchange. Hence, the
coupling (mass flux) between detachment and attachment zones increases with the MinD
recruitment rate, which leads to "canalized transfer".
A recent study [107] has also shown that the interplay between a membrane affinity switch
and cell geometry can lead to an entirely new type of intracellular patterning that is not
based on lateral instabilities (such as the Turing instability) or excitability. In this case the
20
uniform steady state does not become laterally unstable but is replaced by a non-uniform
pattern, i.e. it ceases to exist. The experimental observation [108] is that the MinD homolog
AtMinD from Arabidopsis thaliana forms a bipolar pattern in ∆MinDE E. coli cells. There
is evidence that the ATPase AtMinD can bind (cooperatively) to the membrane in both, its
ADP and its ATP form. Assuming that AtMinD detaches from the membrane in the ADP
bound form, the theoretical analysis shows that the membrane distribution of AtMinD in
steady state is always non-uniform (bi-polar) if, (i) the membrane affinities of AtMinD-ADP
and AtMinD-ADP are different, and (ii) the geometry of the cell deviates from a spherical
shape. The mechanism underlying such geometry induced pattern formation is based on the
local ratio of membrane surface area to cytosolic volume:
in an elliptical cell geometry, a
protein detaching from a cell pole is more likely to re-encounter the membrane in unit time
than a protein that detaches from a site closer to midcell. By utilizing different membrane
affinities (of ADP and ATP states) and cytosolic switching (between these states) a protein
system can then establish highly robust bipolar patterns that reflect the cell's geometry.
Again, they key process underlying pattern formation is cytosolic redistribution – in this
case combined with geometry-dependent reattachment.
MASS-CONSERVING REACTION DIFFUSION MODELS: A NEW PARADIGM?
In the preceding sections we emphasized that mass conservation is the major unifying
property of intracellular pattern-forming protein dynamics. Over the past decade, mass-
conserving reaction-diffusion systems received considerable attention in the theoretical lit-
erature [12, 96, 97, 102, 109–113]. These studies try to answer how mass-conservation affect
reaction-diffusion dynamics. But what is the relevance for biological systems?
In [97] a mass-conserving model for cell polarity is proposed – where mass-conservation
leads to the halting of a propagating wave front. This "pinned" wave represents the polarized
pattern. It has been argued that the corresponding pattern-forming mechanism is not related
to a Turing instability but instead based on excitability and bistability [97, 112]. A similar
line or reasoning is presented in [102]. However, it has been pointed out recently [114, 115]
that a Turing instability in the "wave-pinning" model is recovered upon parameter change.
Other studies report that mass-conserving reaction diffusion systems are prone to coarsening
[96, 109]. However, it remains elusive whether a general relation between mass-conservation
21
and coarsening exists.
Recently, it has been shown that the general mechanism of pattern formation in mass-
conserving reaction diffusion systems is based on the lateral redistribution of the conserved
quantities [12]. The total amount of conserved quantities (protein copy number) determines
the position and stabilities of chemical equilibria. Spatiotemporal redistribution of conserved
quantities shifts local chemical equilibria and is generically induced by any lateral instability
with unequal diffusion coefficients (such as the Turing instability). The pattern forming
dynamics simply follow the movement of local equilibria and the final patterns are scaffolded
by the spatial distribution of local equilibria. This study further demonstrated that "wave-
pinning" patterns originate from the same physical processes as Turing instabilities: the
redistribution of conserved quantities and the shifting local chemical equilibria. In future
research it will be interesting to see how the formation of attachment and detachment zones
can be formalized within the mass-redistribution framework.
SUMMARY AND DISCUSSION
It should be clear by now that activator–inhibitor models do not provide the appropriate
concepts to account for intracellular pattern formation. Rather, the generic key feature of
pattern forming protein system is conformational switching. Proteins cycle between different
states such as active and inactive, or membrane-bound and cytosolic. It is the switching
of (conformational) states that drives the system and leads to pattern formation, not the
production and degradation of proteins, which is the basis of any activator–inhibitor or
activator–depletion model. Any dynamics based on the switching (conformational) states
conserved the total copy number. Therefore, the proper models for intracellular protein
dynamics are mass-conserving reaction-diffusion systems.
In any mass-conserving system pattern formation has to be based on redistribution of
mass. The question to be asked about self-organization in such systems is how redistribution
comes about, i.e. how directed transport is established and maintained.
In the context of intracellular protein pattern formation there is a clear functional division
between membrane-bound and cytosolic protein distributions: The biologically (function-
ally) relevant pattern is that of the membrane-bound factor(s), while the cytosol acts as a
transport medium which facilitates the formation and maintenance of the membrane-bound
22
pattern. The basis of intracellular pattern formation is therefore the emergence of spatially
separated attachment and detachment zones and their coupling (transport from detach-
ment to attachment zone) through cytosolic gradients. The key design principle for pattern
forming protein networks is therefore the ability to set up and maintain attachment and
detachment zones in the presence of ongoing protein redistribution.
We have discussed how pattern formation in three different biological systems is facili-
tated by the formation of attachment and detachment zones. In budding yeast, cell polarity
is established by localized accumulation of Cdc42 on the inner face of the plasma mem-
brane. Pattern formation is based on the localized formation of mutual attachment zones
for Cdc42-GTP and Bem1–GEF complexes through mutual recruitment. In E. coli, pole-
to-pole oscillations emerge due to the interactions of MinD and MinE. This process is based
on the formation of attachment zones with high MinD membrane density due to the re-
cruitment of MinD and MinE from the cytosol. If the balance in a polar zone is shifted
towards higher MinE/MinD ratios, an attachment zone becomes a detachment zone. The
sequestration of MinE in detachment zones enables the formation of new attachment zones
some distance away. The conversion of attachment zones to detachment zones and vice versa
by the slow shift in the MinE/MinD ratio within a zone is the basis for the oscillation. The
establishment of the anterior-posterior axis in C. elegans is based on PAR protein polar-
ization. Here, pattern formation originates from the formation of a (mutual) detachment
zone near midcell, and attachment zones exclusive for pPAR or aPAR, respectively, at the
two cell poles. The establishment and maintenance of this pattern requires that the rates of
both antagonistic processes are balanced.
The question about length scale selection and pattern selection in general is still open.
Apart from mathematically idealized cases that do not apply to biological systems no general
statements about the wavelengths of patterns can be made. However, theoretical studies
of Min protein pattern formation suggest that the emergence and selection of finite wave-
lengths is closely tied to the simultaneous formation and diffusive coupling of attachment
and detachment zones. A key step in the regulation of the diffusive coupling between at-
tachment and detachment zones is the cytosolic switching between conformation with high
and low affinity for the membrane (cytosolic nucleotide exchange). Strikingly, there is evi-
dence that such cytosolic switching processes play a key role in mediating the sensitivity of
self-organized pattern formation to cell geometry [27].
23
Many key questions about intracellular pattern formation remain open.
In our opin-
ion, the focus on concepts based on activator–inhibitor models in the discussion of pattern
formation phenomena has been a hindrance to progress rather than a help.
Intracellular pattern forming protein dynamics are most generally expressed by mass-
conserving reaction-diffusion systems. Local equilibria, as recently suggested [12], are a
promising candidate for a general and unifying theoretical framework for such systems.
To advance our understanding of intracellular protein dynamics a theoretically rigorous
analysis of pattern forming instabilities in mass-conserving reaction-diffusion systems would
be highly desirable.
In his seminal article Turing presented the general idea of pattern
forming instabilities in reaction-diffusion systems. Since its publication 65 years ago only
little has been learned about the general physical principles underlying the Turing instability.
We expect that a focus on the mass-conserving case will finally enable us to extract some
general physical principles of pattern formation systems based on Turing's lateral instability.
ACKNOWLEDGMENTS
This research was funded by the German Excellence Initiative via the NanoSystems Ini-
tiative Munich, and by the Deutsche Forschungsgemeinschaft (DFG) via Project B02 within
SFB 1032 (Nanoagents for SpatioTemporal Control of Molecular and Cellular Reactions),
Areas A and C within GRK2062 (Molecular Principles of Synthetic Biology), and Project
P03 within TRR174 (Spatiotemporal Dynamics of Bacterial Cells). The authors thank
S. Bergeler and W. Daalman for feedback on the manuscript.
∗ [email protected]
[1] Turing AM.
The chemical basis of morphogenesis.
Philosophical Transactions
of the Royal Society of London Series B, Biological Sciences. 1952;237(641):37–72.
DOI: 10.1007/BF02459572.
[2] Cross M, Hohenberg P. Pattern formation outside of equilibrium. Reviews of Modern Physics.
1993;65(3).
[3] Segel La, Jackson JL. Dissipative structure: an explanation and an ecological example.
Journal of theoretical biology. 1972;37:545–559. DOI: 10.1016/0022-5193(72)90090-2.
24
[4] Gierer A, Meinhardt H. A theory of biological pattern formation. Kybernetik. 1972
dec;12(1):30–9.
[5] Meinhardt H, Gierer A. Pattern formation by local self activation and lateral inhibition.
BioEssays. 2000;22:753–760.
[6] Meinhardt H. Models of Biological Pattern Formation: From Elementary Steps to the Or-
ganization of Embryonic Axes. Current Topics in Developmental Biology. 2008;81:1–63.
DOI: 10.1016/S0070-2153(07)81001-5.
[7] Frey E, Halatek J, Kretschmer S, Schwille P. Protein Pattern Formation. Springer-Verlag
GmbH, Heidelberg; 2018. Edited by P. Bassereau and P. C. A. Sens; to be published
[arXiv:1801.01365].
[8] Shamir M, Bar-On Y, Phillips R, Milo R. SnapShot: Timescales in Cell Biology. Cell. 2016
Mar;164(6):1302–1302.e1. DOI: 10.1016/j.cell.2016.02.058.
[9] Kondo S, Miura T. Reaction-diffusion model as a framework for understanding biological
pattern formation. Science. 2010 Sep;329(5999):1616–20. DOI: 10.1126/science.1179047.
[10] Tarnita CE, Bonachela JA, Sheffer E, Guyton JA, Coverdale TC, Long RA, et al. A theoret-
ical foundation for multi-scale regular vegetation patterns. Nature. 2017 01;541(7637):398–
401. DOI: 10.1038/nature20801.
[11] Rietkerk M, van de Koppel J. Regular pattern formation in real ecosystems. Trends Ecol
Evol. 2008 Mar;23(3):169–75. DOI: 10.1016/j.tree.2007.10.013.
[12] Halatek J, Frey E. Rethinking pattern formation in reaction–diffusion systems. Nature
Physics. 2018;DOI: 10.1038/s41567-017-0040-5.
[13] Lutkenhaus J. Assembly dynamics of the bacterial MinCDE system and spatial regula-
tion of the Z ring. Annual Review of Biochemistry. 2007 Jan;76:539–62. DOI: 10.1146/an-
nurev.biochem.75.103004.142652.
[14] Raskin DM, de Boer Pa. Rapid pole-to-pole oscillation of a protein required for directing
division to the middle of Escherichia coli. Proceedings of the National Academy of Sciences.
1999 Apr;96(9):4971–6.
[15] Hu Z, Mukherjee A, Pichoff S, Lutkenhaus J. The MinC component of the division site se-
lection system in Escherichia coli interacts with FtsZ to prevent polymerization. Proceedings
of the National Academy of Sciences. 1999 Dec;96(26):14819–14824.
25
[16] Hu Z, Gogol EP, Lutkenhaus J. Dynamic assembly of MinD on phospholipid vesicles regulated
by ATP and MinE. Proceedings of the National Academy of Sciences. 2002 May;99(10):6761–
6766.
[17] Szeto TH, Rowland SL, Rothfield LI, King GF. Membrane localization of MinD is me-
diated by a C-terminal motif that is conserved across eubacteria, archaea, and chloro-
plasts. Proceedings of the National Academy of Sciences. 2002 Nov;99(24):15693–15698.
DOI: 10.1073/pnas.232590599.
[18] Lackner LL, Raskin DM, de Boer PAJ. ATP-dependent interactions between Escherichia
coli Min proteins and the phospholipid membrane in vitro. Journal of Bacteriology. 2003
Feb;185(3):735–749. DOI: 10.1128/JB.185.3.735-749.2003.
[19] Mileykovskaya E, Fishov I, Fu X, Corbin BD, Margolin W, Dowhan W. Effects of phospho-
lipid composition on MinD-membrane interactions in vitro and in vivo. Journal of Biological
Chemistry. 2003 Jun;278(25):22193–22198. DOI: 10.1074/jbc.M302603200.
[20] Hu Z, Lutkenhaus J. Topological regulation of cell division in Escherichia coli involves rapid
pole to pole oscillation of the division inhibitor MinC under the control of MinD and MinE.
Molecular Microbiology. 1999 Oct;34(1):82–90.
[21] Hu Z, Lutkenhaus J. Topological regulation of cell division in E. coli. spatiotemporal oscilla-
tion of MinD requires stimulation of its ATPase by MinE and phospholipid. Molecular Cell.
2001 Jun;7(6):1337–1343.
[22] Touhami A, Jericho M, Rutenberg AD. Temperature Dependence of MinD Oscillation in
Escherichia coli: Running Hot and Fast. Journal of Bacteriology. 2006 Oct;188(21):7661–
7667. DOI: 10.1128/JB.00911-06.
[23] Park KT, Wu W, Battaile KP, Lovell S, Holyoak T, Lutkenhaus J. The Min Oscillator Uses
MinD-Dependent Conformational Changes in MinE to Spatially Regulate Cytokinesis. Cell.
2011;146(3):396–407.
[24] Schweizer J, Loose M, Bonny M, Kruse K, Monch I, Schwille P. Geometry sensing by
self-organized protein patterns. Proceedings of the National Academy of Sciences. 2012
Sep;109(38):15283–15288. DOI: 10.1073/pnas.1206953109.
[25] Loose M, Fischer-Friedrich E, Ries J, Kruse K, Schwille P. Spatial regulators for bac-
terial cell division self-organize into surface waves in vitro. Science (New York, NY).
2008;320(5877):789–92. DOI: 10.1126/science.1154413.
26
[26] Wu F, van Schie BGC, Keymer JE, Dekker C. Symmetry and scale orient Min protein
patterns in shaped bacterial sculptures. Nature Nanotechnology. 2015 Jun;10:719–726.
[27] Wu F, Halatek J, Reiter M, Kingma E, Frey E, Dekker C. Multistability and dynamic tran-
sitions of intracellular Min protein patterns. Molecular Systems Biology. 2016 jun;12(6):642–
653. DOI: 10.15252/MSB.20156724.
[28] Huang KC, Meir Y, Wingreen NS. Dynamic structures in Escherichia coli: spontaneous
formation of MinE rings and MinD polar zones. Proceedings of the National Academy of
Sciences. 2003 oct;100(22):12724–8. DOI: 10.1073/pnas.2135445100.
[29] Fange D, Elf J. Noise-induced Min phenotypes in E. coli. Plos Computational Biology. 2006
Jun;2(6):e80. DOI: 10.1371/journal.pcbi.0020080.
[30] Halatek J, Frey E. Highly canalized MinD transfer and MinE sequestration explain
the origin of robust MinCDE-protein dynamics. Cell reports. 2012 jun;1(6):741–52.
DOI: 10.1016/j.celrep.2012.04.005.
[31] Hoffmann M, Schwarz US. Oscillations of Min-proteins in micropatterned environ-
ments: a three-dimensional particle-based stochastic simulation approach. Soft Matter.
2014;10(14):2388–2396. DOI: 10.1039/c3sm52251b.
[32] Vecchiarelli AG, Li M, Mizuuchi M, Hwang LC, Seol Y, Neuman KC, et al. Membrane-bound
MinDE complex acts as a toggle switch that drives Min oscillation coupled to cytoplasmic de-
pletion of MinD. Proceedings of the National Academy of Sciences. 2016 Mar;113(11):E1479–
88. DOI: 10.1073/pnas.1600644113.
[33] Halatek J, Frey E.
Effective 2D model does not account for geometry sensing by
self-organized proteins patterns. Proc Natl Acad Sci U S A. 2014 May;111(18):E1817.
DOI: 10.1073/pnas.1220971111.
[34] Kretschmer S, Zieske K, Schwille P. Large-scale modulation of reconstituted Min protein
patterns and gradients by defined mutations in MinE's membrane targeting sequence. PLoS
One. 2017;12(6):e0179582. DOI: 10.1371/journal.pone.0179582.
[35] Fischer-Friedrich E, Meacci G, Lutkenhaus J, Chate H, Kruse K. Intra- and intercellular
fluctuations in Min-protein dynamics decrease with cell length. Proceedings of the National
Academy of Sciences. 2010 Mar;107(14):6134–6139. DOI: 10.1073/pnas.0911708107.
[36] Sliusarenko O, Heinritz J, Emonet T, Jacobs-Wagner C. High-throughput, subpixel precision
analysis of bacterial morphogenesis and intracellular spatio-temporal dynamics. Molecular
27
Microbiology. 2011 Mar;80(3):612–627. DOI: 10.1111/j.1365-2958.2011.07579.x.
[37] Corbin BD, Yu XC, Margolin W. Exploring intracellular space:
function of the Min
system in round-shaped Escherichia coli. The EMBO journal. 2002 Apr;21(8):1998–2008.
DOI: 10.1093/emboj/21.8.1998.
[38] Shih YL, Kawagishi I, Rothfield L. The MreB and Min cytoskeletal-like systems play indepen-
dent roles in prokaryotic polar differentiation. Molecular Microbiology. 2005 Oct;58(4):917–
928. DOI: 10.1111/j.1365-2958.2005.04841.x.
[39] Varma A, Huang KC, Young KD. The Min system as a general cell geometry detection mecha-
nism: branch lengths in Y-shaped Escherichia coli cells affect Min oscillation patterns and di-
vision dynamics. Journal of Bacteriology. 2008 Mar;190(6):2106–17. DOI: 10.1128/JB.00720-
07.
[40] Johnson DI. Cdc42: An essential Rho-type GTPase controlling eukaryotic cell polarity.
Microbiol Mol Biol Rev. 1999 Mar;63(1):54–105.
[41] Chant J, Herskowitz I. Genetic control of bud site selection in yeast by a set of gene products
that constitute a morphogenetic pathway. Cell. 1991 Jun;65(7):1203–12.
[42] Koch G, Tanaka K, Masuda T, Yamochi W, Nonaka H, Takai Y. Association of the Rho
family small GTP-binding proteins with Rho GDP dissociation inhibitor (Rho GDI) in Sac-
charomyces cerevisiae. Oncogene. 1997 Jul;15(4):417–22. DOI: 10.1038/sj.onc.1201194.
[43] Johnson JL, Erickson JW, Cerione RA. New insights into how the Rho guanine nucleotide
dissociation inhibitor regulates the interaction of Cdc42 with membranes. J Biol Chem. 2009
Aug;284(35):23860–71. DOI: 10.1074/jbc.M109.031815.
[44] Bi E, Park HO. Cell polarization and cytokinesis in budding yeast. Genetics. 2012
Jun;191(2):347–87. DOI: 10.1534/genetics.111.132886.
[45] Butty AC, Perrinjaquet N, Petit A, Jaquenoud M, Segall JE, Hofmann K, et al. A positive
feedback loop stabilizes the guanine-nucleotide exchange factor Cdc24 at sites of polarization.
EMBO J. 2002 Apr;21(7):1565–76. DOI: 10.1093/emboj/21.7.1565.
[46] Bose I, Irazoqui JE, Moskow JJ, Bardes ES, Zyla TR, Lew DJ. Assembly of scaffold-mediated
complexes containing Cdc42p, the exchange factor Cdc24p, and the effector Cla4p required
for cell cycle-regulated phosphorylation of Cdc24p. J Biol Chem. 2001 Mar;276(10):7176–86.
[47] Slaughter BD, Das A, Schwartz JW, Rubinstein B, Li R.
Dual modes of
cdc42 recycling fine-tune polarized morphogenesis. Dev Cell. 2009 Dec;17(6):823–35.
28
DOI: 10.1016/j.devcel.2009.10.022.
[48] Kozubowski L, Saito K, Johnson JM, Howell AS, Zyla TR, Lew DJ. Symmetry-breaking
polarization driven by a Cdc42p GEF-PAK complex. Curr Biol. 2008 Nov;18(22):1719–26.
DOI: 10.1016/j.cub.2008.09.060.
[49] Freisinger T, Klunder B, Johnson J, Muller N, Pichler G, Beck G, et al. Establishment of a
robust single axis of cell polarity by coupling multiple positive feedback loops. Nat Commun.
2013;4:1807. DOI: 10.1038/ncomms2795.
[50] Marco E, Wedlich-Soldner R, Li R, Altschuler SJ, Wu LF. Endocytosis optimizes the dynamic
localization of membrane proteins that regulate cortical polarity. Cell. 2007 Apr;129(2):411–
22.
[51] Witte K, Strickland D, Glotzer M. Cell cycle entry triggers a switch between two modes of
Cdc42 activation during yeast polarization. Elife. 2017 Jul;6. DOI: 10.7554/eLife.26722.
[52] Rapali P, Mitteau R, Braun C, Massoni-Laporte A, Unlu C, Bataille L, et al. Scaffold-
mediated gating of Cdc42 signalling flux. Elife. 2017 Mar;6. DOI: 10.7554/eLife.25257.
[53] Woods B, Kuo CC, Wu CF, Zyla TR, Lew DJ. Polarity establishment requires localized
activation of Cdc42. J Cell Biol. 2015 Oct;211(1):19–26. DOI: 10.1083/jcb.201506108.
[54] Irazoqui JE, Gladfelter AS, Lew DJ. Scaffold-mediated symmetry breaking by Cdc42p. Nat
Cell Biol. 2003 Dec;5(12):1062–70. DOI: 10.1038/ncb1068.
[55] Klunder B, Freisinger T, Wedlich-Soldner R, Frey E. GDI-mediated cell polarization in
yeast provides precise spatial and temporal control of Cdc42 signaling. PLoS Comput Biol.
2013;9(12):e1003396. DOI: 10.1371/journal.pcbi.1003396.
[56] Goryachev AB, Pokhilko AV.
Dynamics of Cdc42 network embodies a Turing-
type mechanism of yeast
cell polarity.
FEBS Lett. 2008 Apr;582(10):1437–43.
DOI: 10.1016/j.febslet.2008.03.029.
[57] Wedlich-Soldner R, Altschuler S, Wu L, Li R.
Spontaneous cell polarization through
actomyosin-based delivery of the Cdc42 GTPase.
Science. 2003 Feb;299(5610):1231–5.
DOI: 10.1126/science.1080944.
[58] Wedlich-Soldner R, Wai SC, Schmidt T, Li R. Robust cell polarity is a dynamic state
established by coupling transport and GTPase signaling. J Cell Biol. 2004 Sep;166(6):889–
900. DOI: 10.1083/jcb.200405061.
29
[59] Muller N, Piel M, Calvez V, Voituriez R, Gon¸calves-S´a J, Guo CL, et al. A Predictive
Model for Yeast Cell Polarization in Pheromone Gradients. PLoS Comput Biol. 2016
Apr;12(4):e1004795. DOI: 10.1371/journal.pcbi.1004795.
[60] Hawkins RJ, B´enichou O, Piel M, Voituriez R. Rebuilding cytoskeleton roads: active-
transport-induced polarization of cells. Phys Rev E Stat Nonlin Soft Matter Phys. 2009
Oct;80(4 Pt 1):040903. DOI: 10.1103/PhysRevE.80.040903.
[61] Altschuler SJ, Angenent SB, Wang Y, Wu LF. On the spontaneous emergence of cell polarity.
Nature. 2008 Aug;454(7206):886–9. DOI: 10.1038/nature07119.
[62] Layton AT, Savage NS, Howell AS, Carroll SY, Drubin DG, Lew DJ. Modeling vesicle traffic
reveals unexpected consequences for Cdc42p-mediated polarity establishment. Curr Biol.
2011 Feb;21(3):184–94. DOI: 10.1016/j.cub.2011.01.012.
[63] Savage NS, Layton AT, Lew DJ. Mechanistic mathematical model of polarity in yeast. Mol
Biol Cell. 2012 May;23(10):1998–2013. DOI: 10.1091/mbc.E11-10-0837.
[64] Smith SE, Rubinstein B, Mendes Pinto I, Slaughter BD, Unruh JR, Li R. Independence of
symmetry breaking on Bem1-mediated autocatalytic activation of Cdc42. J Cell Biol. 2013
Sep;202(7):1091–106. DOI: 10.1083/jcb.201304180.
[65] Laan L, Koschwanez JH, Murray AW. Evolutionary adaptation after crippling cell polariza-
tion follows reproducible trajectories. Elife. 2015 Oct;4. DOI: 10.7554/eLife.09638.
[66] Bendez´u FO, Vincenzetti V, Vavylonis D, Wyss R, Vogel H, Martin SG. Spontaneous Cdc42
polarization independent of GDI-mediated extraction and actin-based trafficking. PLoS Biol.
2015 Apr;13(4):e1002097. DOI: 10.1371/journal.pbio.1002097.
[67] Caviston JP, Tcheperegine SE, Bi E. Singularity in budding: a role for the evolutionarily
conserved small GTPase Cdc42p. Proc Natl Acad Sci U S A. 2002 Sep;99(19):12185–90.
DOI: 10.1073/pnas.182370299.
[68] Knaus M, Pelli-Gulli MP, van Drogen F, Springer S, Jaquenoud M, Peter M. Phosphorylation
of Bem2p and Bem3p may contribute to local activation of Cdc42p at bud emergence. EMBO
J. 2007 Oct;26(21):4501–13. DOI: 10.1038/sj.emboj.7601873.
[69] Howell AS, Savage NS, Johnson SA, Bose I, Wagner AW, Zyla TR, et al. Singularity
in polarization:
rewiring yeast cells to make two buds. Cell. 2009 Nov;139(4):731–43.
DOI: 10.1016/j.cell.2009.10.024.
30
[70] Wu CF, Lew DJ. Beyond symmetry-breaking: competition and negative feedback in GTPase
regulation. Trends Cell Biol. 2013 Oct;23(10):476–83. DOI: 10.1016/j.tcb.2013.05.003.
[71] Goehring NW, Trong PK, Bois JS, Chowdhury D, Nicola EM, Hyman AA, et al. Po-
larization of PAR proteins by advective triggering of a pattern-forming system. Science.
2011;334(Nov):1137–1141. arXiv:1011.1669v3. DOI: 10.1126/science.1208619.
[72] Hoege C, Hyman AA. Principles of PAR polarity in Caenorhabditis elegans embryos. Nature
Reviews Molecular Cell Biology. 2013;14(5):315–322. DOI: 10.1038/nrm3558.
[73] Munro E, Nance J, Priess JR. Cortical flows powered by asymmetrical contraction transport
PAR proteins to establish and maintain anterior-posterior polarity in the early C. elegans
embryo. Developmental Cell. 2004 Sep;7(3):413–424. DOI: 10.1016/j.devcel.2004.08.001.
[74] Goehring NW. PAR polarity: From complexity to design principles. Experimental Cell
Research. 2014 Nov;328(2):258–266.
[75] Goehring NW, Trong PK, Bois JS, Chowdhury D, Nicola EM, Hyman AA, et al. Polar-
ization of PAR proteins by advective triggering of a pattern-forming system. Science. 2011
Nov;334(6059):1137–41. DOI: 10.1126/science.1208619.
[76] Meacci G, Ries J, Fischer-Friedrich E, Kahya N, Schwille P, Kruse K. Mobility of Min-proteins
in Escherichia coli measured by fluorescence correlation spectroscopy. Physical biology. 2006
Nov;3(4):255–263. DOI: 10.1088/1478-3975/3/4/003.
[77] Graf IR, Frey E. Generic Transport Mechanisms for Molecular Traffic in Cellular Protrusions.
Physical Review Letters. 2017 Mar;118(12). DOI: 10.1103/physrevlett.118.128101.
[78] Yochelis A, Ebrahim S, Millis B, Cui R, Kachar B, Naoz M, et al. Self-organization of waves
and pulse trains by molecular motors in cellular protrusions. Sci Rep. 2015 Sep;5:13521.
DOI: 10.1038/srep13521.
[79] Pinkoviezky I, Gov NS. Exclusion and Hierarchy of Time Scales Lead to Spatial Segrega-
tion of Molecular Motors in Cellular Protrusions. Phys Rev Lett. 2017 Jan;118(1):018102.
DOI: 10.1103/PhysRevLett.118.018102.
[80] Schumacher D, Bergeler S, Harms A, Vonck J, Huneke-Vogt S, Frey E, et al. The PomXYZ
Proteins Self-Organize on the Bacterial Nucleoid to Stimulate Cell Division. Dev Cell. 2017
May;41(3):299–314.e13. DOI: 10.1016/j.devcel.2017.04.011.
[81] Schumacher D, Søgaard-Andersen L. Regulation of Cell Polarity in Motility and Cell Division
in Myxococcus xanthus. Annu Rev Microbiol. 2017 Sep;71:61–78. DOI: 10.1146/annurev-
31
micro-102215-095415.
[82] Ringgaard S, van Zon J, Howard M, Gerdes K. Movement and equipositioning of plasmids
by ParA filament disassembly. Proc Natl Acad Sci U S A. 2009 Nov;106(46):19369–74.
DOI: 10.1073/pnas.0908347106.
[83] Howard M, Gerdes K. What is the mechanism of ParA-mediated DNA movement? Mol
Microbiol. 2010 Oct;78(1):9–12. DOI: 10.1111/j.1365-2958.2010.07316.x.
[84] Gerdes K, Howard M, Szardenings F. Pushing and Pulling in Prokaryotic DNA Segregation.
Cell. 2010;141(6):927–942. DOI: 10.1016/j.cell.2010.05.033.
[85] Banigan EJ, Gelbart MA, Gitai Z, Wingreen NS, Liu AJ.
Filament depolymeriza-
tion can explain chromosome pulling during bacterial mitosis. PLoS Comput Biol. 2011
Sep;7(9):e1002145. DOI: 10.1371/journal.pcbi.1002145.
[86] Sugawara T, Kaneko K. Chemophoresis as a driving force for intracellular organization:
Theory and application to plasmid partitioning. Biophysics (Nagoya-shi). 2011;7:77–88.
DOI: 10.2142/biophysics.7.77.
[87] Surovtsev IV, Campos M, Jacobs-Wagner C. DNA-relay mechanism is sufficient to explain
ParA-dependent intracellular transport and patterning of single and multiple cargos. Proc
Natl Acad Sci U S A. 2016 Nov;113(46):E7268–E7276. DOI: 10.1073/pnas.1616118113.
[88] Hu L, Vecchiarelli AG, Mizuuchi K, Neuman KC, Liu J. Brownian Ratchet Mechanism for
Faithful Segregation of Low-Copy-Number Plasmids. Biophys J. 2017 Apr;112(7):1489–1502.
DOI: 10.1016/j.bpj.2017.02.039.
[89] Murray SM, Sourjik V. Self-organization and positioning of bacterial protein clusters. Nature
Physics. 2017 Jun;13(10):1006–1013. DOI: 10.1038/nphys4155.
[90] Walter JC, Dorignac J, Lorman V, Rech J, Bouet JY, Nollmann M, et al. Surfing on Protein
Waves: Proteophoresis as a Mechanism for Bacterial Genome Partitioning. Physical Review
Letters. 2017 Jul;119(2). DOI: 10.1103/physrevlett.119.028101.
[91] Ietswaart R, Szardenings F, Gerdes K, Howard M. Competing ParA structures space bac-
terial plasmids equally over the nucleoid. PLoS Comput Biol. 2014 Dec;10(12):e1004009.
DOI: 10.1371/journal.pcbi.1004009.
[92] Bergeler S, Frey E.
Regulation of Pom cluster dynamics in Myxococcus xanthus.
[arXiv:180106133]. 2018;.
32
[93] Huang KC, Ramamurthi KS. Macromolecules that prefer their membranes curvy. Molecular
Microbiology. 2010 Apr;76(4):822–832. DOI: 10.1111/j.1365-2958.2010.07168.x.
[94] Laloux G, Jacobs-Wagner C. How do bacteria localize proteins to the cell pole? Journal of
Cell Science. 2013 Dec;127(1):11–19. DOI: 10.1242/jcs.138628.
[95] Halatek J, Frey E.
Highly canalized MinD transfer and MinE sequestration ex-
plain the origin of robust MinCDE-protein dynamics. Cell Rep. 2012 Jun;1(6):741–52.
DOI: 10.1016/j.celrep.2012.04.005.
[96] Otsuji M, Ishihara S, Co C, Kaibuchi K, Mochizuki A, Kuroda S. A mass conserved reaction-
diffusion system captures properties of cell polarity. PLoS Comput Biol. 2007 Jun;3(6):e108.
DOI: 10.1371/journal.pcbi.0030108.
[97] Mori Y, Jilkine A, Edelstein-Keshet L. Wave-pinning and cell polarity from a bistable
reaction-diffusion system. Biophysical journal. 2008 may;94(9):3684–97. DOI: 10.1529/bio-
physj.107.120824.
[98] Howell AS, Jin M, Wu CF, Zyla TR, Elston TC, Lew DJ. Negative feedback enhances
robustness in the yeast polarity establishment circuit. Cell. 2012 Apr;149(2):322–33.
DOI: 10.1016/j.cell.2012.03.012.
[99] Das M, Drake T, Wiley DJ, Buchwald P, Vavylonis D, Verde F. Oscillatory dynamics
of Cdc42 GTPase in the control of polarized growth. Science. 2012 Jul;337(6091):239–43.
DOI: 10.1126/science.1218377.
[100] Yang L, Dolnik M, Zhabotinsky AM, Epstein IR. Turing patterns beyond hexagons and
stripes. Chaos: An Interdisciplinary Journal of Nonlinear Science. 2006;16(3):037114.
http://aip.scitation.org/doi/pdf/10.1063/1.2214167. DOI: 10.1063/1.2214167.
[101] Otsuji M, Ishihara S, Co C, Kaibuchi K, Mochizuki A, Kuroda S. A mass conserved reaction-
diffusion system captures properties of cell polarity. PLoS Comput Biol. 2007 Jun;3(6):e108.
[102] Semplice M, Veglio A, Naldi G, Serini G, Gamba A. A bistable model of cell polarity. PLoS
One. 2012;7(2):e30977. DOI: 10.1371/journal.pone.0030977.
[103] Berry J, Brangwynne C, Haataja MP. Physical Principles of Intracellular Organization via
Active and Passive Phase Transitions. Reports on Progress in Physics. 2018 Jan;.
[104] Yang HC, Pon LA. Actin cable dynamics in budding yeast. Proc Natl Acad Sci U S A. 2002
Jan;99(2):751–6. DOI: 10.1073/pnas.022462899.
33
[105] Maini PK, Woolley TE, Baker RE, Gaffney EA, Lee SS. Turing's model for biological
pattern formation and the robustness problem.
Interface Focus. 2012 Aug;2(4):487–96.
DOI: 10.1098/rsfs.2011.0113.
[106] Levine H, Rappel WJ. Membrane-bound Turing patterns. Phys Rev E. 2005 Dec;72(6 Pt
1):061912.
[107] Thalmeier D, Halatek J, Frey E. Geometry-induced protein pattern formation. Proceedings
of the National Academy of Sciences. 2016;113(3):548–553. DOI: 10.1073/pnas.1515191113.
[108] Zhang M, Hu Y, Jia J, Gao H, He Y. A plant MinD homologue rescues Escherichia coli
HL1 mutant (DeltaMinDE) in the absence of MinE. BMC Microbiol. 2009 May;9:101.
DOI: 10.1186/1471-2180-9-101.
[109] Ishihara S, Otsuji M, Mochizuki A. Transient and steady state of mass-conserved reaction-
diffusion systems. Phys Rev E Stat Nonlin Soft Matter Phys. 2007 Jan;75(1 Pt 2):015203.
DOI: 10.1103/PhysRevE.75.015203.
[110] Pogan A, Scheel A. Layers in the Presence of Conservation Laws. J Dyn Diff Equat. 2012
Apr;24:249–287.
[111] Goh RN, Mesuro S, Scheel A. Spatial Wavenumber Selection in Recurrent Precipitation.
SIAM Journal on Applied Mathematics. 2011;10(1):360–402.
[112] Mori Y, Jilkine A, Edelstein-Keshet L. Asymptotic and bifurcation analysis of wave-pinning
in a reaction-diffusion model for cell polarization. SIAM Journal on Applied Mathematics.
2011;71(4):1401.
[113] Kessler DA, Levine H. Nonlinear self-adapting wave patterns. New Journal of Physics. 2016
Dec;18(12):122001. DOI: 10.1088/1367-2630/18/12/122001.
[114] Goryachev AB, Leda M. Many roads to symmetry breaking: molecular mechanisms
and theoretical models of yeast cell polarity. Mol Biol Cell. 2017 Feb;28(3):370–380.
DOI: 10.1091/mbc.E16-10-0739.
[115] Trong PK, Nicola EM, Goehring NW, Kumar KV, Grill SW. Parameter-space topology of
models for cell polarity. New Journal of Physics. 2014;16.
34
FIG. 1. Turing's general linear stability analysis and heuristic ad-hoc interpretations in terms of
the activator–inhibitor picture based on the production and degradation of reactants. (a) Any
random perturbation (black line) of a spatially uniform state (gray line) can be decomposed into
Fourier modes. Linear stability analysis yields the growth rates σq of the amplitude of all modes
q. This is represented in the dispersion relation (blue line σq). Unstable modes (marked red) grow
in amplitude an determine the pattern emerging out of the random perturbation during the incip-
ient time evolution. (b) The activator–inhibitor model is based on autocatalytic production of a
slowly diffusing activator, which in turn stimulates the production of a fast-diffusing inhibitor that
suppresses autocatalytic activator production. Both activator and inhibitor are subject to degra-
dation. The faster diffusion of the inhibitor leads to the formation of an inhibition zone in which
degradation dominates over activator (and inhibitor) production. (c) In the activator–depletion
model the inhibitor is replaced by a substrate that is subject to degradation, and autocatalytic
activator production is replaced by the autocatalytic conversion of substrate into activator. The
rate of conversion is limited by the available substrate. Heuristically, this conversion could be
equated with the attachment of cytosolic proteins to the membrane. However, the reverse process
(detachment) is not taken into account. Both substrate and activator are steadily degraded and are
produced at a finite rate. If the activator density is too low, the conversion process is suppressed
and the degradation process dominates, as in the activator–inhibitor model.
35
=+++++Conc.Spaceautocatalyticproductionproductiondominatesdegradationdominatesslow activator diffusionfast inhibitor diffusionActivator-Inhibitorbautocatalyticconversionconversiondominatesdegradationdominatesfast substrate diffusionslow activator diffusionActivator-Depletioncno detachmentatimeevolutionSpaceConc.band ofunstable modesFIG. 2. Biochemical interaction networks of three model systems for self-organized intracellular
pattern formation. (a) The Min system of E. coli [28, 30]. (b) Cdc42 system of S. Cerevisiae
[49, 55] .
(c) PAR system of C. elegans [71].
(d) Switching between two conformal states of
the proteins involved is a recurring theme in the biochemical networks (a-c). Cycling between
membrane bound and cytosolic states is driven by the ATPase/GTPase cycle of MinD and Cdc42
respectively, while the PAP-proteins each cycle between different phosphorylation states. In general
we expect switching between distinct conformal states – catalyzed by "switch regulators" such as
NTPase-activating proteins (NAPs), nucleotide exchange factors (NEFs), phophatases, and kinases
– to be a core element of biochemical networks that mediate intracellular pattern formation.
36
Bem1-GEF mediatedCdc42 recuritmentand activationCdc42-GDPGDIBem1GEFCdc42-GTP recruits Bem1-GEFmembranecytosolCdc42 GTPase cycleMinD-ATPMinD-ADPMinEmutual detachmentof aPAR and pPARaPARpPARMinD recruitsitself and MinEMinDE-complexdetachmentacbDTNEFNAPNDPNTPPDPhosphataseKinasePhosphorylatedDephosphorylateddABSwitch regulatorsState AState BFIG. 3. Linear (in)stability of a uniform initial distribution of proteins. (a) In a uniform steady
state, attachment and detachment must balance everywhere. An external cue or a random per-
turbation due to stochastic noise can lead a local increase in membrane density. How the relative
balance of attachment and detachment processes shifts in the region of increased membrane den-
sity, determines the stability of the uniform state. If the balance in a region of increased membrane
density shifts in favour of attachment (b), the region becomes an attachment zone leading to a
further increase in membrane density due to redistribution through the cytosol. Hence the spatially
separated attachment and detachment zones are maintained, leading the establishment of a pat-
tern. If the balance in a region of increased membrane density shifts the attachment-detachment
balance in favour of detachment (b'), this region becomes a detachment zone, while the region of
lower membrane density becomes an attachment zone, such that the system returns to its uniform
balanced state (c').
37
attachment zonedetachment zoneCytosol concentrationrandom perturbationattachment zonedetachment zoneredistributionReturn to uniform stateredistributionIncrease of membrane density shifts reaction balancein favour of attachementattachment zonedetachment zoneredistributionmembranediffusionuniform attachment-detachment balanceMaintained pattern stateTime evolutionIncrease of membrane density shifts reaction balancein favour of detachementabb'cc'FIG. 4. Mutual recruitment of Cdc42 and Bem1–GEF complexes is the core mechanism of Cdc42
polarization. (a) The interaction network of Cdc42, Bem1 and GEF (Cdc24) can conceptually
be simplified to two key processes: Bem1–GEF complexes on the membrane recruit Cdc42-GDP
from the cytosol, which is followed by immediate nucleotide exchange (conversion to Cdc42-GTP).
Reversely, Cdc42-GTP recruits Bem1 to the membrane by, where it immediately recruits Cdc24 to
form Bem1–GEF complexes. (b) A local accumulation of Cdc42-GTP thereby acts as recruitment
template for Bem1–GEF complexes (1), creating an attachment zone for Bem1–GEF, while de-
tachment of Bem1–GEF dominates in zones of low Cdc42-GTP density (2). Bem1–GEF complexes
then accumulate in their attachment zone. This accumulation acts as a recruitment template for
Cdc42 (2), creating co-polarized attachment zones of Cdc42 and Bem1–GEF (1') and (2'). (c)
Taken together, the mutual recruitment processes establish and maintain Cdc42-polarization.
38
++mutual recruitmentzonedetachment zoneredistribution++cytosolic densitycytosolic densityBem1-GEFrecruitmentBem1,GEFdetachmentredistributionCdc42 recruitment,nucleotide exchangeCdc42 detachmentredistribution++Bem1-GEF template for Cdc42 recruitmentCdc42 template for Bem1-GEF recruitmentBem1-GEFCdc42-GDPMutual recruitment ofBem1-GEF and Cdc42-GTPabc11'22'FIG. 5. Pole-to-pole oscillations of MinD and MinE: MinD recruits both itself and MinE from the
cytosol creating an attachment zone for both proteins (1). As MinDE complexes accumulate in the
polar zone, their detachment begins to dominate over MinD attachment. (2) The old polar zone
traps MinE because it is both an attachment and a detachment zone for MinE, which only cycles
locally, as long as there is free MinD left on the membrane. This allows cytosolic MinD to form a
new polar zone at the other end of the cell. MinE trapping ends when all MinD has detached from
the old polar zone, such that the new polar zone becomes an attachment zone for MinE (3), and
the process starts over at the opposite end of the cell (4).
39
+cytosolic density+redistributionMinDE detachmentMinD recruits itselfand MinE++MinD self-recruitmentMinDE detachmentlocal MinE cyclingredistribution++MinD self-recruitmentMinDE detachmentlocal MinE cyclingredistribution++redistributionMinDE detachmentMinD recruits itselfand MinE1234FIG. 6. Mutual antagonism between aPAR and pPar proteins creates a detachment zone at the
interface between aPAR- and pPAR-dominated regions. A region of high aPAR density on the
membrane is a detachment zone for pPAR, such that detaching pPARs can only attach to a region
of low aPAR density, and vice versa. A balance of the mutual antagonistic processes is necessary to
prevent one protein species from dominating over the other and taking over the whole membrane.
40
mutual detachmentaPAR attachmentpPAR attachmentredistributionredistributioncytosolic densitiesFIG. 7. Canalized transfer of MinD imposes a length scale intrinsic to the reaction–diffusion
dynamics in the model developed by Halatek and Frey [30]. (a) Canalized transfer refers to the
case where attachment-flux and detachment-flux are of similar magnitude such that the cytosolic
density does not vary (left). To maintain the flux between detachment and attachment zone a
cytosolic gradient must be maintained. The length scale of this gradient dictates the distance
between attachment and detachment zones independently of system size. Thus, canalized transfer
can give rise to patterns with an intrinsic wavelength in large systems (right). (b) If a detachment
zone forms when no attachment zone is available, the detaching proteins fill up a cytosolic reservoir.
Once the reservoir reaches a critical density an attachment zone will form. In this case no length
scale between attachment and detachment zones is set, since both zones exchange proteins through
a (uniform) cytosolic reservoir.
41
Simultaneous attachment and detachmentSequential attachment and detachmentweak attachmentstrong detachmentcytosolic reservoirstrong attachmentno detachmentcytosolic reservoirLength = LLength = 2Lno attachmentdetachmentattachment zonedetachment zonemembranecytosolcanalized transferintrinsic transportlength scaleweak attachmentstrong detachmentcytosolic reservoirstrong attachmentno detachmentcytosolic reservoirno attachmentdetachmentTime evolutionattachment zonedetachment zonecanalized transferdetachment zonecanalized transferintrinsic transportlength scaleintrinsic transportlength scaleabCytosolic reservoir does not impose length scaleCanalized transfer determines pattern length scale |
1001.4850 | 1 | 1001 | 2010-01-27T04:35:43 | Multi Scale Investigation of Surface Topography of Ball Python (Python Regius) Shed Skin in Comparison to Human skin | [
"physics.bio-ph"
] | Constructing a surface that is an integral part of the function of tribosystems (deterministic surface) is an intriguing goal. Inspirations for such surfaces come from studying natural systems and deducing design rules. The major attraction is that natural systems, while functionally complex, are, in general, of optimized shape and performance. It is further believed that functional complexity of natural systems is what affords natural species to morph continuously to adapt with the operating environment. One bio-species that is of interest is the Ball Python. This is because such a species continuously slides against various surfaces, many of which are deemed tribologically hostile, without sustaining much damage. Much of the success of that species in adapting to its environment is attributed to surface design features. In that respect, studying these features and how do they contribute to the control of friction and wear is very attractive. This paper is a step in this direction. In this work we apply a multi scale surface characterization approach to study surface design features of the Python Regius. The focus is on those features that are typically used to assess the performance of high quality lubricating surfaces. To this end, topographical features are studied by SEM and through White Light Interferrometery (WLI). We probe the roughness of the surface on multi scale and as a function of location within the body. In addition we draw a comparison of these features to those of human skin. | physics.bio-ph | physics | Multi Scale Investigation of Surface Topography of Ball Python (Python Regius) Shed Skin
in Comparison to Human skin
H. A. Abdel-Aal 1 M. El Mansori S. Mezghani
Arts et Métier ParisTech, Rue Saint Dominique BP 508,
51006 Chalons-en-Champagne,
France
1 corresponding author: [email protected]
ABSTRACT
A major concern in designing tribo-systems is to minimize friction, save energy, and to reduce
wear. Conventional philosophy for design centers on mechanical and material considerations. In
particular designers pay more attention to material properties and material choices based on
mechanical properties rather than the design and shape of the contacting surfaces and the relation
of that surface to the function of the device. As a result of thriving for miniaturization, focus has
shifted toward optimal surface design (that is to construct a surface that is an integral part of the
function of the tribosystem). Inspirations for such a trend come from studying natural systems
and mimicking natural design rules. The major attraction is that natural systems, while
functionally complex, are, in general, of optimized shape and performance. It is further believed
that functional complexity of natural systems is what affords natural species to morph
continuously to adapt with the operating environment. One bio-species that is of interest is the
Ball Python. This is because such a species continuously slides against various surfaces, many
of which are deemed tribologically hostile, without sustaining much damage. Much of the
success of that species in adapting to its environment is attributed to surface design features. In
that respect, studying these features and how do they contribute to the control of friction and
wear is very attractive. This paper is a step in this direction. In this work we apply a multi scale
surface characterization approach to study surface design features of the Python Regius. The
focus is on those features that are typically used to assess the performance of high quality
lubricating surfaces (such as those obtained through plateau honing). To this end, topographical
features are studied by SEM and through White Light Interferrometery (WLI). We probe the
roughness of the surface on multi scale and as a function of location within the body. In addition
we draw a comparison of these features to those of human skin.
1. Introduction
In order for many next generation products to succeed they must offer greater functiona lity and
improved performance. Ultra precision and structured surfaces are increasingly being adopted to
gain such advantages. As such, the development of higher performing products through the
application of ultra precise, complex and structured surfaces, such as those textured through
plateau honing for improved lubrication capability, is an active area of research. In seeking
inspirations for such custom designed surfaces many engineers turn toward natural systems (i.e.,
bio-species, plants, insects etc.,). There are many features that deserve attention within natural
systems. These may include superior functionality, the ability to harness functional complexity to
achieve optimal performance, and harmony between shape form and function. From a tribology
perspective, the existence of surfaces that are design features, of the particular species, intended
to facilitate functional performance is a point of deserving interest. Tribological investigations
often deal with complex systems that, while nominally homogeneous, are practically
1
compositionally heterogeneous. Compositional heterogeneity is either inherent (structural), or
evolutionary (functional) [1]. Inherent heterogeneity is due to initial variation in composition,
material selection, component chemistry, etc. Evolutionary heterogeneity, on the other hand,
arises because of the evolution of the local response of different parts of a sliding assembly
during operation. Subsystem components, for example, since they entertain different loads will
react in a manner that is proportional to the local loading conditions. Distinct responses cause
system subcomponents to evolve into entities that differ from their initial state. System
heterogeneity, thus, introduces a level of functional complexity to the sliding assembly.
Functional complexity, in turn, characterizes the interaction of system subcomponents, and of the
system as a unit, with the surroundings. Most of such interaction, it is to be noted, takes place
through the surface. Natural systems, regardless of the degree of functional complexity inherited
within, display harmonious characteristics and an ability to self-regulate. Whence, as a general
rule they operate at an optimal state marked by economy-of-effort. Man-engineered systems
(MES), in contrast, do not exhibit such a level of optimized performance.
Much of the ability of natural systems to self regulate is attributed to optimized relationship
between shape, form and function especially when surface design is considered. That is, shape
and form in natural systems are always targeted toward optimal function. Such customization,
however, is not advanced in MES. Bio-species, in that respect, offer many a lesson. This is
because biological materials, through million years of existence, have evolved optimized
topological features that enhance wear and friction resistance [2]. One species of remarkable
tribological performance that may serve as an inspiration for optimal surface texturing is that of
snakes. This is due to the Objective-targeted design features associated with their mode of
legless locomotion.
Snakes are limbless animals. They have multi-modes of motion (slithering, crawling, serpentine
movement etc) that take place during propulsion. Such motion modes are initiated through
muscular activity. That is, through a sequence of contraction and relaxation of appropriate
muscle groups. Transfer of motion between the body of the snake and the substrate depends on
generation of sufficient tractions. This process, generation of traction and accommodation of
motion, is handled through the skin. Thus the skin of the snake assumes the role of motion
transfer and accommodation of energy consumed during the initiation of motion.
The number, type and sequence of muscular groups responsible for the initiation of motion, and
thus-employed in propulsion, will vary according to the particular mode of motion to be initiated.
It will also depend on the habitat and the surrounding environment. This also will affect the
effort invested in initiation of motion and thereby also affects the function of the different parts
of the skin and the amount of energy required to be accommodated. So that, in general, different
parts of the skin will have different functional requirements. Moreover, the life habits of the
particular species, e.g., defense, hunting, and swallowing) will require different deterministic
functions of the different parts of the skin. That is the snake species, is a true representative of a
heterogeneous tribo-system with a high degree of functional complexity, despite which, they
don’t suffer damaging levels of wear and tear.
Many researchers investigated the intriguing features of the serpentine family. Adam and Grace
[3] studied the ultra structure of pit organ epidermis in Boid snakes to understand infrared
2
sensing mechanisms. Johnna et al [4] investigated the permeability of shed skin of pythons
(python molurus, Elaphe obsolete) to determine the suitability as a human skin analogue.
Mechanical behavior of snake skin was also a subject of several studies as well. Jayne [5]
examined the loading curves of six different species in uni-axial extension. His measurements
revealed substantial variation in loads and maximum stiffness among samples from different
dorsoventral regions within an individual and among homologous samples from different
species. Rivera et al.[6] measured the mechanical properties of the integument of the common
garter snake (Thomnophis sirtalis-Serpentine Colubridae). They examined mechanical properties
of the skin along the body axis. Data collected revealed significant differences in mechanical
properties among regions of the body. In particular, and consistent with the demands of
macrophagy, it was found that the pre-pyloric skin is more compliant than post pyloric skin.
Prompted by needs to design bio-inspired robots several researchers probed the frictional
features of snake motion to understand the mechanisms responsible for regulating legless
locomotion. Hazel et al [7] used AFM scanning to probe the nano-scale design features of three
snake species. The studies of Hazel and Grace revealed the asymmetric features of the skin
ornamentation to which both authors attributed frictional anisotropy.
In order to mimic the beneficial performance features of the skin, an engineer should be provided
with parametric guidelines to aid with the objective-oriented design process. These should not
only be dimensional. Rather, they should extend to include metrological parameters used to
characterize tribological performance of surfaces within the MES domain. Thus, in order to
deduce design rules there exists a need for quantification of the relationship governing micro -
structure and strength topology of the bio-surface; exploring the quantitative regulation of macro
and micro texture, and finally devising working formulae that describe (and potentially predict)
load carrying capacity during locomotion in relation to geometrical configuration at both the
micro and the macro scale. This paper is a preliminary step toward that goal.
In this work, we apply a multi-scale surface characterization approach to probe the design
features of shed skins obtained from a Ball Python (Python Regius). Such a species is of
tribological interest because of its locomotion taking place within a non-breakable boundary
lubrication regime. Such a performance feature is facilitated by the topology of skin
ornamentation. This renders such a species of interest to designers of plateau honing engineers
where surfaces are designed for minimal lubricant consumption and for designers of hip and
knee prostheses where maintaining a continuous boundary lubrication regime is a must. The
emphasis in the current work, therefore, is on deducing those metrological aspects of the shed
skin that are deemed essential to quantify the quality of tribological performance of industrial
sliding assemblies (e.g., cylinder-piston).
2. Background
2.1. The Python species
Python Regius, figure 1-a, is a constrictor type non-venomous snake species that is typically
found in Africa. Due to the reaction of the animal of curling into a spherical position, where the
head is tucked under the trunk, figure 1-b,, for protection when distressed, it is also called “ball-
python”. The build of the snake is non-uniform as the head-neck region as well as that of the tail
is thinner than the-region containing the trunk. The trunk meanwhile is the region of the body
where most of the snake body mass is concentrated. Consequently it is more compact and thicker
3
than other parts with the overall cross section of the body is more elliptical rather than circular.
The tail section also is rather conical in shape. Skin of the python contains blotches imposed
upon an otherwise black background. The blotches are of a non uniform shape and contour and
are colored in dark and light brown. The ventral part of the body is typically cream, or extremely
light yellow, color with occasional black spots scattered within.
.
Figure 1. A Ball Python (Python Regius) species., a) normal position, b) curled-up “Ball”
position
2.2.
general features
The skin of reptiles, as well as that of snakes, is covered with scales. These maybe thought of as
elevated folds or wrinkles within the skin. The pattern of the scales, shape, number, and
ornamentation differ according to the species. A snake will typically hatch with a constant
number of scales. This number remains fixed throughout the life duration of the snake. Each
snake type has a different number of scales that is unique to the particular species. While the
total number of scales remains constant, the size and shape change during the life span of the
snake to accommodate growth and changes in size.
There are several functions for scales in a snake. They provide protection from the environment,
they regulate energy exchange between the animal and environment, further they aid in
camouflage and capture of prey [8]. Scales also perform a tribological function as they regulate
frictional interaction, and help generate tractions upon propulsion. In snakes, skin cell division,
occur periodically and that causes the reptile, to grow a second skin underneath the current (old)
skin, and then “sheds” the old one. That is due to the cell division process the snake periodically
sheds its entire outer skin layer.
3. Materials and Methods
All observations reported herein pertain to shed skin obtained from a 115 cm, 14 years old male
Ball Python (Python Regius) housed individually in a glass container with news paper subs trate.
For optical microscopy observations skin was observed as is, whereas all samples for Scan
Electron Microscopy observations were coated by a vacuum deposited Tungsten layer of
thickness 10 nm. Surface topography analysis took place through two methods: SEM imaging in
topography mode and through examination using a white light interferometer (WYKO NT3300
3D Automated Optical Profiling System). For comparison purposes, we obtained replicas of
human skin. Samples were obtained by replicating the skin of a 40 years old female in two
places: back of the hand and upper inside portion of the arm. All replicas were made using a light
4
silicone rubber impression material (Silflo ®-Lexico, Plandent, UK) after using alcohol to clean
subject skin.
4. Results and discussion
4.1.
Initial Observations
Initial observations on the structure of the scales were performed using two methods:
photography of the life species and optical microscopy. These were done without any treatment
of the skin.
4.2. Photographical Observations
Figure 2 (a and b) depicts details of the surface structure of two regions on the life snake. Figure
2-a details the skin geometry in the head region from the inner (sliding side), whereas figure 2 -b
depicts the details of the stomach (belly-again the sliding side) of the snake. The photographs
reveal that polygons constitute the geometrical building block of the surface. This polygon has
eight sides, octagon, in the general area of the mouth (represented by the letter O in the figure).
Past a line that joins the eyes, line AA, the pattern of the skin changes to hexagonal. The size of
individual unit cells (octagons or hexagons) is not generally the same. The size of the octagons
in the mouth region is not uniform. However, compared to the hexagonal patterns within the
throat region, the area of the unit octagon is, in general, greater than that of the hexagonal unit.
Hexagonal cells on the other hand are of uniform shape and size and seem to be of uniform
density per unit area.
Within the mouth zone (above the line AA) the aspect ratio of the octagonal unit cell is,
qualitatively, uniform. The major axis of the polygon, moreover, appears to point in the same
direction of the body major longitudinal axis of the snake (BMLA). Considering the function of
the head frontal part, feeding, some design features may be pointed out. Pythons feed by
constricting the prey then swallowing it to be digested in the stomach. Upon inhaling the prey,
the jaws have to stretch to accommodate the volume, and shape of the p rey. The prey may offer
some resistance that leads to multi-axial displacements of the surface material. Under such
conditions, the design requirement of the surface is to allow for global flexibility and to facilitate
local multi-axis displacements (stretching) all while minimizing possible damage to the skin.
Observing the build of the skin above line AA, one would note that the skin is built of small
patches of uniform shape (octagon), and are linked by a channel of what seems to be flexible
strips of skin. Such a design allows for multi-axis stretching of the surface with possible damping
of sudden jerks (caused by prey resistance) provided through the flexible links. A similar trend
is noted in the throat region. Here the need to accommodate volume dilatation is greater since
the prey by this point would, most likely, be subdued. Within this zone hexagonal unit cells
make up the surface.
5
body major axisAAOHO-OctagonH-HexagonBody Longitudinal major axis unit cell of ventral skin building block dorsal scalesventral scalesFigure 2. Details of scale structure at two positions on the life species.
a) Shields within the head-throat region
b) Ventral scales close to the waist section.
The uniform distribution of hexagonal cells is likely to aid the compliance of the surface and
increase its flexibility. Interestingly such a hexagonal pattern is noted to be the most efficient
way to pack the largest number of similar objects in a minimum space [9] . Naturally, had the
skin been made of larger patches, or one continuous sheet, the probability of surface rupture
would have increased. Figure 2-b depicts the surface geometry in the sliding side of the skin
(general region of the belly). The photograph reveals that the hexagonal pattern constitutes the
basic building block of the skin. The size of the hexagons differs around the circumference of
the body. Large cells are particular to the main sliding area whereas cells of smaller size are
particular to the back and the sides. Consistent with the construction of the head-throat region,
the aspect ratio of the cells is variable. Of interest is the orientation of the major axis of the skin
unit cells with respect to the BMLA. In the head-throat region the major axis of the cells is
oriented parallel to the body major axis whereas in the belly (ventral scales) the major cell axis is
perpendicular to the main body axis. Varenberg and Gorb [10], based on experiments on the
hexagonal structures found on tarsal attachment pads of the bush cricket (tettigonia viridissima)
suggest that variation in the aspect ratio of hexagonal structures may alter the friction force of
elastomers by at least a factor of two. Additionally, we propose that the perpendicular orientation
of the cells, with respect to the major axis of the snake, within the main sliding region aids in
shifting the weight, and hence the contact angle and area of the snake upon sliding. Note that
since the body of the snake is of cylindrical shape, the highest curvature of the skin will be
oriented along the major cell axis. As such, upon sliding, the area of contact, and therefore the
total tractions, will depend on the direction of motion (higher sideways and minimal forward).
That is the orientation of the hexagon axis renders the friction forces isotropic. Such an
observation is consistent with the findings of Zhang et al [11] who studied the frictional
mechanism and anisotropy of Burmese python's ventral scales. They reported that the friction
coefficient of the ventral scale had closely relationship with moving direction. The frictional
coefficient for backward and lateral motion was one third higher than that in forward motion.
4.3. Structure of Shed Skin
4.3.1. Optical Microscopy Observations
In snakes the epidermis is made up of different layers with the innermost called the stratum
germinativum. The outer layers, which are renewed during shedding, are, from the inside, -,
mesos-, -layer, and Oberhautchen. The Oberhautchen, mainly consisting of -keratin, is in
direct contact with the environment and possesses a fine surface structure called micro-
ornamentation [12]. Details of the micro-ornamentation were described by earlier authors [13-
16]. Initial observations on the structure of the scales were performed using optical microscopy.
without any treatment of the skin. Figure 3 depicts the structure of the scales at two positions
within the skin in a region close to the waist of the snake. The first was from the back (dorsal
scale) whereas the second position represented the belly of the snake (ventral). Note that
although the general form of the cells is quite similar for both positions the size of a unit cell
within the skin is quite different in both cases. In particular the cell is wider for the ventral
(belly) positions. Each cell (scale) is also composed of a boundary and a membrane like
structure. Note also the overlapping geometry of the skin and the scales (the so called scale and
6
hinge structure). The skin from the inner surface hinges back and forms a free area, which
overlaps the base of the next scale, which emerges below this scale figure 4.
Figure 3. The structure of the scales on the inside of the shed skin at a region close to the mid
section of the species at two orientations: back (dorsal) and abdominal (ventral).
Figure 4. The details of dorsal scales from the inside of shed skin. The terminology used is:
membrane to denote the major area of the scale and boundary to denote the raised part forming
the circumference of the scale. Scale marker is 1mm.
4.3.2. Scan Electron Microscopy (SEM) Observations
Four positions on the shed skin were identified for initial examination. The choice of the
positions was based on the functional profile of each position in the live species (figure 5).
Position I is a representative of the neck region, position II represents the beginning of the trunk
(waist) region, position III marks the boundary between the trunk and the tail region, finally
position IV represents the tail region (containing the so called subcaudal scales). Skin swatches
from each of the chosen positions were examined at different magnifications (X250-X15000) in
topography mode. In order to suppress charging phenomena and improve the quality of
observation, the surface of each sample was metalized by depositing a 10nm thin layer of
platinum (Pt) using a sputter coater EMITECH K575X. Comparison of the uncoated and coated
pictures is given in figure 6.
7
membraneBoundaryHinge
Figure 5. Equivalent positions chosen on the snake shed skin for SEM observations
Figure 6: Comparison of the quality of SEM pictures at two positions on the shed skin at
X1000. Upper row represents pictures of the uncoated skin samples while the lower row
represents those of the coated samples.
8
IIIIIIIVDark Skinlight skinFor each position samples from the dark and the light colored skin (see figure 5) were also
examined along with samples from the underside of the body. Major features of the observations
are shown in figures 7 and 8.
ventral scales- Scale boundary
ventral scales: Membrane
dorsal scales- light skin boundary
dorsal scales- light skin Membrane
dorsal scales- dark skin boundary
dorsal scales- dark skin Membrane
Figure 7: Major features of SEM observations of the skin swatches.(X-250)
9
ventral scales- Scale boundary
ventral scales: Membrane
dorsal scales- light skin boundary
dorsal scales- light skin Membrane
dorsal scales- dark skin boundary
dorsal scales- dark skin Membrane
Figure 8: Major features of SEM observations of the skin swatches.(X-5000)
Note the inner structure that comprises pores. Two types of pores (or micro pits) may be
distinguished: those located within the boundary and those located within the membrane. Image
analysis of the pictures indicates that the diameter of the boundary-pores ranges between (200
nm – 250 nm). An estimate of the diameter of the membrane-pores was estimated by Hazel et al
10
[7] using AFM analysis to be in the range (50nm -75 nm). Worth noting, is that within the same
study [7] Hazel and co-workers scanned the contour of the fibril tips, within the membrane,
using AFM and deduced that the tips are of asymmetric shape. Based on this observation,
combined to the inter-lamellar layout of the fibrils, they suggested that such design features
would lead to anisotropic frictional properties of the body. The locomotion mechanism
responsible for such anisotropy is, however, considered, beyond the scope of the current work
since the focus is on the topographical features and the metrology of the surface.
Surface protrusions are also noted within the boundary. These protrusions are of an asymmetric
shape and irregular distribution. The surface of the membrane also comprises micro-nano fibrile
structures. These are not of consistent shape and spacing. Note for example that the shape of
fibril located in the dark colored skin region is different than that located within the light colored
skin region (compare the X-5000 pictures). Moreover the density of the fibrils seems to be
different within the different color regions (denser within the dark colored region). The spacing
between different rows of fibrils also differs by skin color and position within the body (ventral>
dorsal light skin> dorsal dark skin). Further analysis of images revealed that the density of the
boundary-pores vary by position. That is the number of pores per unit area is not constant along
the body, rather it changes relative to the position. Figure 9 is a plot of the variation in the
density of the pores relative to the two sides of the skin (back-Dorsal scales and abdominal-
Ventral scales) and in relation to the color of the skin (Light Patches (L) Vs Dark Patches D)
within the back also.
Figure 9. The variation in the density of the boundary pores (pore /mm2) with position, and with
color of skin.
4.4. Metrology of the Surface
Examination of the surface topography features of the skin using White Light Interferometry
(WLI) on a swatches of skin (1500 μm x 1500 μm) yielded the basic parameters that describe the
surface (asperity radii and curvature etc., ). Figure 10 depicts a typical WLI graph of the skin.
The shown inteferogram pertains to a skin spot that is located along the waist of the snake from
11
02004006008001000IIIIIIVIVentralDorsal-lightDorsal-DarkPosition on BodyDensity of Boundary Pores (pore/mm)2the belly side (ventral). Two interferograms are depicted: the one to the right hand side of the
figure represents the topography of the cell- membrane whereas the one depicted to the left
represents a multi-scale scan for the whole skin swatch. Note the scale on the right of the pictures
as it indicates the deepest valley and highest point of the skin topography. For these typical skin
swatches the value of the deepest part of the membrane was about -120 μm, whereas the highest
summit is about 100 μm. The comparable values for the whole swatch are about -5.5 μm and 8.2
μm respectively.
Figure 10. Multi scale WLI graphs depicting the topography of the skin building block (Scale)
boundary and membrane (scale bar is in m).
To establish a measure of comparison between the topography of the snake skin, we compared
scans of snake skin to human skin scans. The results are given in figure 11.
Human scans from two different positions are depicted in the figure. The first scan is from skin
located at the back of the hand whereas the second pertains to skin located at the inside of the
upper arm of a 40 years old Caucasian female. Samples were obtained by replicating the skin
using a replicating silicone. For the snake we chose to compare scans of the ventral scales
located in the waist section of the body since it is a major load bearing area during locomotion.
Again, it is noted that the peak summit and valley values of the surface of the snake skin are in
the order of magnitude of one third that of human skin. Such a comparison highlights,
qualitatively, the origins of the superior tribological performance exhibited by the snake. Note
that under the conditions of examination in this work quantitative comparison between both skin
types are not possible. This is because the tribological performance of human skin is not a mere
function of surface topography. Rather, it also depends on the water content within skin cells.
As such, comparing the frictional features of snake skin used in the current work, which is dry by
default since it is shed from the species, to human skin in vivo is meaningless. Moreover, the
friction of skin in general is dominated by acoustic emission. Again, provisions to monitor such
effect were not undertaken in this study as it was judged that such feature warrants detailed
investigation (which is currently undergone in our laboratory).
12
Human skin-back of the hand
Human skin-inside of upper arm
Python- Ventral close to waist-membrane
Figure 11. Comparison between the surface texture of snake skin and that of human skin
replicas as revealed by White light interferometer (scale bar in m)
4.5. Bearing curve analysis
To complete the analysis, we studied the load bearing characteristics of the skin at each of
the predetermined positions (I through IV). Surface parameters were extracted from SEM
topography photographs. The complete set of analyzed pictures provided a matrix of roughness
parameters that describe the texture of the shed skin at variable scales ranging from X-100 to X-
5000. Table 1 (a and b) provides a summary of the parameters extracted from the analysis. It
can be seen that the scale of the analysis affects the value of the parameters, which may point at a
fractal nature of the surface. Discussion of the implication of such finding is considered out of
the scope of this work. However, of interest is to point out one of the features that directly relates
to the design of the surface. Comparing the ratios between the Reduced Peak Height Rpk, Core
Roughness Depth Rk, and Reduced Valley Depth Rvk reveals symmetry between the positions
(compare the columns Rpk/Rk, Rvk/Rk, and Rvk/Rpk of table 1-b, and figure 12 a and b). This
symmetry is interesting on the count that positions II and III represent the boundaries of the main
load bearing regions (trunk). That is the regions on the body where the snake has most of his
body weight concentrated (refer to figure 5) and thereby it is the region that is principally used in
locomotion. Such symmetry may very well be related to the wear resistance ability of the
surface or to the boundary lubrication quality of locomotion. Of interest also, is to find if
13
implementing a surface of such characteristic parameters (functionally textured surface) in
plateau honing for example would be conducive to an anti-scuffing and economical lubricant
consumption performance. Such a point is a subject of ongoing investigation.
Table-1 effect of magnification on surface parameters
a- Surface parameters based on X-250 pictures
Position I
Position II
Position III
Position VI
Cr/Cf Cl/Cf Rpk/Rk
0.718
0.861
0.391
0.612
2.010
2.011
0.733
1.628
2.066
1.388
0.926
0.617
Rvk/Rk
0.159
0.545
0.436
0.195
Rvk/Rpk
1.144
0.656
0.621
0.749
b- Surface parameters based on X-5000 pictures
Position I
Position II
Position III
Position VI
Cr/Cf Cl/Cf Rpk/Rk
0.654
1.207
1.930
0.478
1.803
1.273
0.484
1.622
1.671
1.772
1.158
0.636
Rvk/Rpk
0.679
0.812
0.800
0.688
Rvk/Rk
0.285
0.404
0.359
0.260
Figure 12: Plot of the ratio of the load bearing parameters Rvk/Rk and Rpk/Rk at two
magnifications X-250 and X-5000.
5. Conclusions and Future Outlook
In this work we presented the results of an initial study to probe the geometric features of the
skin of the Python Regius. It was found the structure of the unit cells is of regionally similar
shape (octagonal and hexagonal).
Although almost identical in size and density, the skin constituents (pore density and essential
size of the unit cell) vary by position on the body. Comparison of the topography of the snake
skin to that of a human female revealed that the surface roughness of the snake species is around
one third that of the human samples.
Analysis of the surface roughness parameters implied a multi-scale dependency of the
parameters. This may point at a fractal nature of the surface a proposition that needs future
verification. The analysis of bearing curve characteristics revealed symmetry between the front
14
Rpk/RkR/RvKk00.10.20.30.40.50.6IIIIIIIVX250X5000Position on body00.10.20.30.40.50.60.70.8IIIIIIIVX250X5000R/RpKkPosition on bodyand back sections of the body. It also revealed that the trunk region is bounded by two cross-
sections of identical bearing curve ratios. This has implications in design of textured surfaces
that retain an unbreakable boundary lubrication quality and high wear resistance.
Clearly much work is needed to further probe the essential features of the surface geometry.
Namely, the basic parametric make up of the topography, form, and their relation to friction and
wear resistance.
Acknowledgment
The authors acknowledge Mr. Benjamin Favre for assistance in SEM imaging and Mrs Ruth Ann
Jones, of Troup county GA School System, for donating the shed skin used in this work.
References
1. Abdel-Aal, H. A., Complexity, Synergy and emergence in tribosystems, Int. J. Mat. Prod.
Tech, 38 (1), 1-3, (2010).
2. Tong, J., H. Wang, Ma, Y.Ren, L.: Two-body abrasive wear of the outside shell surfaces
of mollusc Lamprotula fibrosa Heude, Rapana venosa Valenciennes and Dosinia anus
Philippi, Trib. Let. 19, 331-338 (2005).
3. Adam B. Safer and Michael S. Grace, Infrared imaging in vipers: differential responses
of crotaline and viperine snakes to paired thermal targets, Behavioral Brain Research 154,
1, 55-61 (2004)
4. Johnna B. Roberts, Harvey B. Lillywhite, Lipids and the permeability of epidermis from
snake, Journal of Experimental Zoology 228, 1 - 9 (2005).
5. Jayne, B.C. : Mechanical behavior of snake skin. J. Zool. 214, 125-140 (1988)
6. G. Rivera, A. H. Savitzky, J. A. Hinkley, Mechanical properties of the integument of the
common gartersnake, Thamnophis sirtalis (Serpentes: Colubridae), J. Exp. Bio. 208,
2913-2922 (2005).
7. J. Hazel, M. Stone, M. S. Grace and V. V. Tsukruk. : Nanoscale design of snake skin for
reptation locomotion via friction anisotropy, Journal of Biomechanics 32, 477-484
(1999).
8. Sepkoski, J.: A compendium of fossil marine animal genera (Reptilia entry). Bulletins of
American Paleontology 364, 560 (2002).
9. Ball, P., The Self-Made Tapestry: Pattern Formation in Nature, Oxford University Press,
New York (2001)
10. Varenberg, M. and S. N. Gorb. :Hexagonal surface micropattern for dry and wet friction,
Advanced Materials 21, 483-486 (2009).
11. Zhang, H., Z. Dai, Yang, S.:Structure and friction characteristics of snake abdomen,
Journal of Nanjing University of Aeronautics and Astronautics 40, 360-363 (2008).
12. Price R.M.: Microdermatoglyphics: an appeal for standardization of methodology and
terminology with comments on recent studies of North American natricines, J Herpet.
24, 324–325(1990)
13. R. Ruibal.: The ultrastructure of the surface of lizard scales, Copeia 4, 698–703 (1968)
14. Price R. M. , Kelly P.: Microdermatoglyphics: basal patterns and transition zones, J
Herpet. 23,244–261(1989)
15. Chiasson R.B., Bentley D.L., Lowe C.H.: Scale morphology in Agkistrodon and closely
related crotaline genera. Herpetologica 45, 430–438 (1989)
15
16. Gower D. J.: Scale micro-ornamentation of uropeltid snakes, J Morph. 258, 249–268
(2003).
16
|
1006.1825 | 2 | 1006 | 2010-12-03T19:31:41 | Kinetic Theory of Flocking: Derivation of Hydrodynamic Equations | [
"physics.bio-ph",
"cond-mat.stat-mech"
] | It is shown how to explicitly coarse-grain the microscopic dynamics of the Vicsek model for self-propelled agents. The macroscopic transport equations are derived by means of an Enskog-type kinetic theory. Expressions for all transport coefficients at large particle speed are given. The phase transition from a disordered to a flocking state is studied numerically and analytically. | physics.bio-ph | physics | Kinetic Theory of Flocking: Derivation of Hydrodynamic Equations
Thomas Ihle1, 2
1Department of Physics, North Dakota State University, Fargo, North Dakota, 58108-6050, USA
2Max Planck Institute for the Physics of Complex Systems, Nothnitzer Strasse 38, 01187 Dresden, Germany
It is shown how to explicitly coarse-grain the microscopic dynamics of the rule-based Vicsek
model for self-propelled agents. The hydrodynamic equations are derived by means of an Enskog-
type kinetic theory. Expressions for all transport coefficients at large particle speed are given. The
phase transition from a disordered to a flocking state is studied numerically and analytically.
0
1
0
2
c
e
D
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
5
2
8
1
.
6
0
0
1
:
v
i
X
r
a
Pattern formation and collective motion in systems of
self-propelled objects are fascinating phenomena which
have attracted much attention. Systems of interest in-
clude animal flocks [1], chemically powered nanorods [2],
and actin networks driven by molecular motors [3]. Theo-
retical studies of these systems are usually based on phe-
nomenological transport equations.
In most cases, the
equations are postulated by means of symmetry argu-
ments, which define only the general form of the terms
but leave their coefficients undetermined.
One goal of this Letter is to provide a system-
atic derivation of all relevant coefficients for the two-
dimensional Vicsek model (VM) of self-propelled parti-
cles [4]. In the VM, pointlike particles are driven with
constant speed. At each time step, a given particle as-
sumes the average direction of motion of its neighboring
particles, with some added noise. As the noise amplitude
decreases, the system undergoes a phase transition from
a disordered state, in which the particles have no pref-
ered global direction, to an ordered state, in which the
particles move collectively in the same direction. This
long-range order motivated renormalization group stud-
ies by Toner and Tu [5]. They found that the stabiliza-
tion of the ordered phase is due to the nonzero speed
of the particles, allowing two originally distant particles
to interact with each other at a later time. The phase
transition was originally thought to be continuous [4] but
recent numerical work [6] indicates that the transition is
discontinuous with strong finite size effects. There are
few analytical studies on this transition [7, 8]. They do
not treat the original VM but simple models related to
it. For example, Bertin et al.
[7], study a model with
simplified interactions and a continuous time dynamics
by means of a Boltzmann equation.
Numerical simulations of the VM [4, 6] show localized
high-density structures, for which a Boltzmann descrip-
tion, which is restricted to low densities, is not sufficient.
Enskog's proposal to generalize the Boltzmann equation
to dense gases was a major milestone in kinetic theory. In
this Letter, it is shown how an Enskog-type equation with
genuine multi-body collisions can be obtained for the VM
and how this can be used to rigorously derive hydrody-
namic equations. In addition to the terms postulated by
Toner and Tu [5], the derived equations contain several
new relevant terms which describe an intricate coupling
between density and order parameter gradients. The co-
efficients of all terms, compatible with the symmetries
of the system, are calculated explicitly in third order of
a gradient expansion. The new kinetic equation is used
to determine the mean-field phase diagram of the VM,
which agrees well with direct numerical simulations but
disagrees with the results of a related continuous time
model [7]. This shows the importance of explicitely tak-
ing the discrete time, rule-based nature of the VM into
account. The derived hydrodynamic equations are ap-
plied to study the stability of a homogeneous flocking
state against spatio-temporal perturbations.
I discuss
how an instability at the onset of collective motion can
change the appearance of the phase transition from sec-
ond to first order. Predictions for the system size where
this change is expected to happen, are given.
In the VM, a system of N pointlike particles with
continuous spatial coordinates ri(t) and velocities vi(t)
evolves via two steps: streaming and collision. During
a time step τ , particles stream ballistically: xi(t + τ ) =
xi(t) + τ vi(t). The magnitude of the particle velocities is
fixed to v0. Only the directions θi of the velocity vectors
are updated in the collision step: a circle of radius R is
drawn around a given particle and the average direction
¯θi of motion of the particles within the circle is deter-
mined according to ¯θi = arctan[Pn
j cos(θj )].
The new directions follow as θi(t + τ ) = ¯θi(t) + ξi, where
ξi is a random number chosen with uniform probabil-
ity from the interval [−η/2, η/2]. Since explicitly coarse-
graining the dynamics of the VM is difficult, in previ-
ous work [9], I have first validated the formalism on a
simpler equilibrium model [10] which shares essential fea-
tures with the VM. The kinetic formalism starts with the
Liouville equation for the N-particle probability density
j sin(θj)/Pn
P (θ(N ), X(N ) + τ V(N ), t + τ ) =
×Z 2π
0
dθ(N ) P (θ(N ), X(N ), t)
N
Yi=1
1
dξ(N )
−η/2
ηN Z η/2
δ(θi − ξi − ¯θi)
(1)
where X(N ) ≡ (x1, x2, . . . , xN ), θ(N ) ≡ (θ1, θ2, . . . , θN ),
and δ(x) = P∞
m=−∞ δ(x + 2πm) is the periodically
The velocities V(N ) ≡
continued delta function.
(v1, v2, ..., vN ), are given in terms of angle variables,
vi = v0(cos θi, sin θi). The collision integral contains
integrations over the pre-collisional angles θj. Assum-
ing that the particles are uncorrelated prior to the col-
lisions, the probability distribution can be expressed as
a product of identical one-particle probability distribu-
i=1 P1(θi, xi). This approxi-
mation of molecular chaos is valid at moderate and large
noise strength η and when the mean free path (mfp) is
large compared to the radius of interaction R. Here, the
mfp is defined as the distance a particle travels between
collisions, τ v0, and is density-independent due to the dis-
crete nature of the dynamics. Multiplying Eq. (1) by
tions: P (θ(N ), X(N )) = QN
Pi δ(v − vi)δ(x − xi) and integrating over all particle
positions xi and angles θi, yields in the large N -limit [9],
a kinetic equation for the one-particle distribution func-
tion, f (θ, x, t) = N P1(θ, x, t),
4
η
3
C
2
1
Numerics: Nagy et al.
Theory: Aldana et al.
Numerics
0
1
2
a) M
3
4
5
2
0.01
ω
R
L*
103
102
101
0.01
100
1
M
0
0
b) k
0.2
0.4
0.6
FIG. 1. a) The critical noise ηC as a function of the average
number of collision partners, M = ρ0πR2, and the prediction
of Eq. (35) for large v0 from Ref. [7], (dashed line), in com-
parison with results from Refs. [4, 8, 11]. b) Real part of the
growth rate, ωR, of a small longitudinal perturbation of the
ordered state versus dimensionless wave number k at M = 5,
very close to the threshold, (ηC −η)/ηC = 0.00057. The insert
shows a lower and an upper bound for the crossover length
L∗ (in units of the mfp) beyond which the phase transition is
expected to become discontinuous.
f (θ, x + τ v, t + τ ) =
×f (θ1, x, t) δ(θ − ξ − ¯θi)
n
Yi=2
1
η Z η/2
−η/2
e−MR
dξ(cid:28)(cid:28) N
Xn=1
f (θi, xi, t)(cid:29)θ(cid:29)x
n!
n
(2)
ability. Below a critical noise ηC (ρ0) there exists another
fixed-point solution which breaks rotational symmetry.
It has a maximum at some arbitrary angle θ and de-
scribes ordered motion into this direction. The critical
noise follows from the condition λ = 1, with
0
where MR(x, t) = RR ρ(y, t) dy is the average number
of particles in a circle of radius R centered around x.
The local particle density ρ is given as a moment of the
distribution function, ρ(x, t) = R 2π
0 f (θ, x, t) dθ; h...ix =
RR ... dx2 dx3...dxn denotes the integration over all posi-
tions, n − 1 particles can assume within the interaction
circle; h...iθ = R 2π
...dθ1dθ2...dθn is the average over all
pre-collisional angles of n particles in the interaction cir-
cle. Since particles in the VM have zero volume, there
is a non-zero probability that a large number of parti-
cles can be found in the collision circle of a given parti-
cle. This leads to the unusual structure of the collision
integral in which every term in the sum accounts for a
n-particle collision. For example, the n = 4 term involves
the product of four distribution functions and describes a
four body collision. Interactions between particles which
are not at the same position but a distance ≤ R apart
are explicitely taken into account by Eq. (2). This leads
to collisional momentum transfer which is a key feature
of the Enskog equation and not included in Boltzmann-
type equations. Hence, Eq. (2), can be interpreted as an
Enskog-like equation for pointlike particles with discrete
time evolution; it remains valid even at infinite density.
Let us first consider a spatially homogeneous system
and study stationary solutions of Eq. (2). This amounts
to solving the fixed-point equation f0(θ) = C(f0) for the
stationary distribution function f0, where C denotes the
r.h.s. of Eq. (2). It can be easily checked that the con-
stant distribution f0 = ρ0/2π is a fixed-point at any noise
and average density, ρ0 = N/A, where A is the area of
the system. This solution corresponds to the disordered
phase, where all velocity directions occur at equal prob-
λ =
I(n) =
2(cid:17) e−MR
4
η
sin(cid:16) η
(2π)n Z 2π
1
0
N
Xn=1
n2M n−1
R
n!
I(n)
dθ1 . . .Z 2π
0
dθn cos ¯θ cos θ1
(3)
Here, MR is equal to πR2ρ0 and ¯θ is the average angle de-
fined above Eq. (1). The fixed-point equation was solved
numerically for η ≤ ηC . The solution approaches a cosine
with vanishing amplitude when η approaches the critical
noise. By means of a Fourier cosine series in θ − θ the
behavior at the critical point was extracted analytically.
The order parameter, defined as the amplitude g1 of the
first non-trivial Fourier coefficient, is found to behave as
g1 ∝ √ηC − η. Thus, the order-disorder transition ap-
pears to be continuous with the mean-field critical expo-
nent of 1/2. Fig. 1a) shows the calculated phase diagram
(solid line). Evaluating Eq. (3) in the low density limit
gives ηC ∝ R√ρ. This scaling with the square root of the
density agrees with previous numerical [4] and theoreti-
cal results [6, 7]. However, there is no dependence of the
critical noise on the particle speed in the large mfp limit,
which is consistent with numerical simulations of the VM
[4, 11] but disagrees with the scaling ηC ∝ √ρRv0 for
[7]. The dashed
ρ → 0 of the continuous model of Ref.
line in Fig 1a) shows that the phase diagram of this model
(obtained from Eq. (35) in [7] with v0τ /R = 5) does not
describe the VM. Evaluating Eq. (3) in the infinite den-
sity limit yields ηC → 2π. In order to see whether the
homogeneous ordered state is stable under time evolu-
tion, I derive the hydrodynamic equations by means of a
Chapman-Enskog expansion [9, 12]. The basic idea be-
hind this expansion is to take the local stationary state
as a reference state and expand around it in powers of
the hydrodynamic gradients. To systematically account
for these gradients a dimensionless ordering parameter ǫ
is introduced, which is set to unity at the end of the cal-
culation. The procedure starts with a Taylor expansion
of the l.h.s of Eq. (2) around (θ, x, t). The spatial gradi-
ents that occur are scaled as ∂α → ǫ∂α, and multiple time
scales ti are introduced in the temporal gradients. These
time scales describe different physical processes, for ex-
ample, in regular fluids, the time scale proportional to ǫ
describes convection. For the VM, this is expressed as
∂t = ∂t0 + ǫ∂t1 + ǫ2∂t2 . . . .
Expanding the distribution function and the collision
integral in powers of ǫ, f = f0 + ǫf1 + ǫ2f2 + ..., and
C = C0 + ǫC1 + ǫ2C2 + ..., inserting into Eq. (2), and
collecting terms of the same order in ǫ leads to a hi-
erarchy of evolution equations for the fi. Due to the
absence of momentum conservation and Galilean invari-
ance this set of equations is dramatically different from
the usual one.
It is not a priori evident whether the
scaling ansatz for the time derivatives is correct. How-
ever,
it turns out that this choice avoids any incon-
sistencies if additionally the expansion of the distribu-
tion function f is identified as an angular Fourier series
with f0(x, t) = ρ(x, t)/2π and, for n > 0, fn(x, θ, t) =
[an(x, t) cos (nθ) + bn(x, t) sin (nθ)] /πvn
0 .
Many moments of the collision integral such as
hvxvyC2i = R 2π
0 vxvyC2 dθ are required in the Chapman-
Enskog expansion. For simplicity, these moments are
evaluated in the limit of large mfp, τ v0 ≫ R. This in-
volves solving the following four integrals,
Jm(n) =
1
(2π)n Z 2π
0
dθ1 . . .Z 2π
0
dθn Ψm
(4)
where Ψm is given by Ψ1 = cos2 ¯θ cos 2θ1, Ψ2 =
cos ¯θ sin ¯θ cos θ1 sin θ2, Ψ3 = cos ¯θ cos θ1 cos 2θ2, and
Ψ4 = cos ¯θ cos θ1 cos θ2 cos θ3. The average angle ¯θ is
a function of the angles θ1, θ2, . . . θn.
moments of f , namely the particle density ρ = R 2π
(wx, wy), ~w = R 2π
We seek a hydrodynamic description of the first two
0 f dθ
and the macroscopic momentum density vector ~w =
0 ~vf dθ. Inserting the Fourier represen-
tation of f into these moments shows that the first order
coefficients are given by the momentum density, a1 = wx
and b1 = wy. Multiplying the hierarchy of evolution
equations by powers of the microscopic velocity vector
~v = (vx, vy) and integrating over θ gives a set of equa-
tions for the time development of the density and the
moments ai and bi. This analysis is performed in the
vicinity of the critical point, λ − 1 ≪ 1, in order to sig-
nificantly simplify the consistent closure of the hierarchy
of moment equations, see [11].
For simplicity, all equations are rescaled by expressing
time in units of τ and distances in units of the mfp, τ v0,
which also makes ρ and ~w dimensionless. After straight-
3
j
1
hj
1+p
8(p−1)
qj
S
kj
S
2(p−1)
8(p−1)
96(p−1)2
2 − p2+10p+1
3 − q
2(p−1)
4
q(1+p)
4(p−1)2
4(p−1)2
− S
Γ − Sq
2 − Sq(p−3)
2(p−1)2
p−1
Γ
5 − q(p2+10p+1)
48(p−1)3
Γ
24 − Sq(p2
−2p+13)
24(p−1)3
Γ
4(p−1)
96(p−1)2
− S(p+5)
4 − Sq
12 − Sq(p−4)
− Sq(p+5)
48(p−1)3
12(p−1)2
Γ
TABLE I. The transport coefficients hj , qj and kj, defined in
Eq. (6), are expressed as functions of Γ, S, p, q, see Eq. (8).
forward, but tedious, calculations one obtains the con-
tinuity equation ∂tρ + ∂αwα = 0, and a rotationally-
invariant equation for the momentum density,
∂t ~w +∇ · H = −b∇ρ + (λ− 1) ~w + Q1 · ~w + Q2 ·∇ρ (5)
with b = (3 − λ)/4. The momentum flux tensor H and
the tensors Q1, Q2,
5
5
5
H =
hi Ωi
Q1 =
qi Ωi Q2 =
ki Ωi
Xi=1
Xi=1
Xi=1
(6)
are given in terms of five symmetric traceless tensors Ωi,
Ω1,αβ = ∂αwβ + ∂βwα − δαβ∂γwγ
Ω2,αβ = 2∂α∂βρ − δαβ∂2
γρ
Ω3,αβ = 2wαwβ − δαβw2
Ω4,αβ = wα∂βρ + wβ ∂αρ − δαβwγ∂γρ
Ω5,αβ = 2(∂αρ)(∂βρ) − δαβ(∂γ)2 .
(7)
The tensor Ω1 is the viscous stress tensor of a two-
dimensional fluid. The transport coefficients in Eq. (6)
are given in Table I. They depend on the following vari-
ables,
p =
4
η
sin (η)
N
Xn=1
e−MR
n!
n2M n−1
R J1(n)
q =
S =
4πγ2
η
8πγ2
η
sin (η)
e−MR
n!
n2(n − 1)M n−2
R J2(n)
sin
η
2
e−MR
n!
n2(n − 1)M n−2
R J3(n)
(8)
N
N
Xn=2
Xn=2
Xn=3
N
Γ =
8π2γ4
3η
sin
η
2
e−MR
n!
n2(n − 1)(n − 2)M n−3
R J4(n)
where γ is the ratio of the interaction radius to the mfp,
γ = R/τ v0. Eq. (5) is consistent with the one postulated
in Ref. [5] but contains additional gradient terms. It has
a homogeneous flocking solution: ~w = w0 n and ρ = ρ0.
The amplitude of the flow is given by w0 = p(1 − λ)/q3.
In order to study the spontaneous onset of collective
motion, a perturbation around this state is considered,
ρ(x, t) = ρ0 + δρ eik·x+ωt, ~w(x, t) = w0 n + δ ~w eik·x+ωt,
and Eq. (5) is linearized in δρ and δ ~w. The characteristic
equation for the growth rate ω(~k) describes three possible
modes. I found that in a small window, ηS < η < ηC , di-
rectly below the onset of flocking, one of the longitudinal
modes is always unstable against long wavelength pertur-
bations: the real part of ω is positive for 0 < k < k0 as
shown in Fig. 1b). A similar instability was reported by
Bertin et al. [7]. Chat´e et al. [6] found numerically that
the order/disorder transition is discontinuous for system
sizes L larger than the crossover length L∗. Assuming
that the long wave instability is the reason for this fi-
nite size effect, I calculated the largest value of k0 within
the narrow instability window at constant density, k∗, in
order to obtain a lower bound for L∗. Plotting 2π/k∗
gives the lower curve in the insert of Fig. 1b). An up-
per bound was obtained by determining the wave number
kmax where the growth rate has the largest value inside
the instability window. The upper curve in the insert
shows 2π/kmax as a function of density. The minimum
around M ≈ 2 and the divergences at small and large
densities are consistent with numerical results [6].
To see what happens to a growing perturbation be-
yond the linear instability, the continuity equation and
Eq. (5) were integrated on a L × L lattice with peri-
odic boundaries by means of a predictor-corrector scheme
[13]. These simulations confirmed that the ordered
phase is stable for small system sizes L < 2π/k0. For
slightly larger system sizes one observes a stable, inho-
mogeneous steady state with a global order parameter,
h ~wi = R ~w dx/L2, larger than the amplitude of the homo-
geneous state, w0. Finally, for much larger system sizes,
it turns out that the system is both linearly and nonlin-
early unstable for ηS < η < ηC . Longitudinal pertur-
bations grow without bound; they do not lead to stable
solitons as suggested in Ref. [7]. However, direct simula-
tions of the VM at large mfp do show solitary structures
such as traveling high-density bands in a window just be-
low the transition [6, 11]. At lower noise these structures
disappear.
Identifying this "solitary" window with the
instability window, its size can be predicted by the cur-
rent theory which takes all the details of the VM such
as multi-body interactions into account. However, inside
this window, the hydrodynamic equations are driven out
of the range of their validity and are not suited to describe
solitons. Nagy et al [4] did not see high-density bands at
small velocity v0. To treat this limit of small mfp theoret-
ically, one has to abandon the molecular chaos approx-
imation i.e. go beyond the mean-field approximation,
which is outside the scope of this paper.
4
In summary, a first-principle derivation of the hydro-
dynamic equations of the VM by means of a novel ki-
netic theory is presented and a stability analysis of the
resulting equations, Eq. (5), is performed. The mean-
field phase diagram for arbitrary density is calculated. It
agrees within a few percent with simulation results and is
shown to be independent of the particle speed in the large
mfp limit. It is also shown that the continuous theory of
[7] fails to reproduce the phase diagram of the VM and
that one has to explicitely incorporate the discrete time
dynamics and genuine multi-body interactions in order
to achieve agreement. The theory presented here is con-
sistent with numerical studies [4, 6], and suggests the fol-
lowing picture of the nature of the flocking transition in
the large mfp limit considered here: At η = ηC a homo-
geneous ordered state bifurcates continuously from the
disordered state. At the threshold, this state is unstable
to longitudinal, long wavelength fluctuations. Perturba-
tions from a large range of wave numbers k < k0 become
unstable, already in close vicinity to the threshold. The
transition appears to be continuous in small systems but
becomes a discontinuos transition in large systems due
to the emergence of density waves which abruptly in-
crease the global order parameter. An estimate of the
system size L∗, above which the discontinuous nature of
the transition is expected to emerge, is given. This length
is found to diverge at small and large densities, consistent
with numerical results.
Support from the National Science Foundation under
grant No. DMR-0706017 is gratefully acknowledged. I
thank F. Julicher, L.S. Schulman, H. Chat´e, A. Denton,
A. Wagner and D. Kroll for valuable discussions.
[1] I.D. Couzin et al., Nature 433, 513 (2005).
[2] Y.G. Tao and R. Kapral, Soft Matter 6, 756 (2010).
[3] J. F. Joanny et al., New J. Phys. 9 422 (2007).
[4] T. Vicsek et al., Phys. Rev. Lett. 75, 1226 (1995); A.
Czir´ok, H. E. Stanley, T. Vicsek, J. Phys. A, 30, 1375
(1997); M. Nagy, I. Daruka, T. Vicsek, Physica A 373,
445 (2007).
[5] J. Toner and Y. Tu, Phys. Rev. E 58, 4828 (1998).
[6] G. Gr´egoire and H. Chat´e, Phys. Rev. Lett. 92, 025702
(2004); H. Chat´e et al., Phys. Rev. E 77, 046113 (2008);
Phys. Rev. Lett. 99, 229601 (2007).
[7] E. Bertin, M. Droz, and G. Gr´egoire, Phys. Rev. E 74,
022101 (2006); J. Phys. A 42, 445001 (2009).
[8] M. Aldana et al., Phys. Rev. Lett. 98, 095702 (2007).
[9] T. Ihle, Phys. Chem. Chem. Phys. 11, 9667 (2009).
[10] G. Gompper et al., Adv. Polym. Sci. 221, 1 (2009).
[11] T. Ihle, A. Gebremariam, in preparation.
[12] D. McQuarrie, Statistical Mechanics, New York, 1976.
[13] R.W. MacCormack, AIAA Journal 20, 1275 (1982).
|
1803.09516 | 1 | 1803 | 2018-03-26T11:24:30 | Role of salt valency in the switch of H-NS proteins between DNA-bridging and DNA-stiffening modes | [
"physics.bio-ph",
"q-bio.BM",
"q-bio.GN"
] | This work investigates the interactions of H-NS proteins and bacterial genomic DNA through computer simulations performed with a coarse-grained model. The model was developed specifically to study the switch of H-NS proteins from the DNA-stiffening to the DNA-bridging mode, which has been observed repeatedly upon addition of multivalent cations to the buffer, but is still not understood. Unravelling the corresponding mechanism is all the more crucial, as the regulation properties of H-NS proteins, as well as other nucleoid proteins, are linked to their DNA-binding properties. The simulations reported here support a mechanism, according to which the primary role of multivalent cations consists in decreasing the strength of H-NS/DNA interactions compared to H-NS/H-NS interactions, with the latter ones becoming energetically favored with respect to the former ones above a certain threshold of the effective valency of the cations of the buffer. Below the threshold, H-NS dimers form filaments, which stretch along the DNA molecule but are quite inefficient in bridging genomically distant DNA sites (DNA-stiffening mode). In contrast, just above the threshold, H-NS dimers form 3-dimensional clusters, which are able to connect DNA sites that are distant from the genomic point of view (DNA-bridging mode). The model provides clear rationales for the experimental observations that the switch between the two modes is a threshold effect and that the ability of H-NS dimers to form higher order oligomers is crucial for their bridging capabilities. | physics.bio-ph | physics | Role of salt valency in the switch of H-NS proteins
between DNA-bridging and DNA-stiffening modes
Marc Joyeux(#)
Laboratoire Interdisciplinaire de Physique,
CNRS and Université Grenoble Alpes,
Grenoble, France
Running title: bridging/stiffening modes of H-NS
Abstract: This work investigates the interactions of H-NS proteins and bacterial genomic
DNA through computer simulations performed with a coarse-grained model. The model was
developed specifically to study the switch of H-NS proteins from the DNA-stiffening to the
DNA-bridging mode, which has been observed repeatedly upon addition of multivalent
cations to the buffer, but is still not understood. Unravelling the corresponding mechanism is
all the more crucial, as the regulation properties of H-NS proteins, as well as other nucleoid
proteins, are linked to their DNA-binding properties. The simulations reported here support a
mechanism, according to which the primary role of multivalent cations consists in decreasing
the strength of H-NS/DNA interactions compared to H-NS/H-NS interactions, with the latter
ones becoming energetically favored with respect to the former ones above a certain threshold
of the effective valency of the cations of the buffer. Below the threshold, H-NS dimers form
filaments, which stretch along the DNA molecule but are quite inefficient in bridging
genomically distant DNA sites (DNA-stiffening mode). In contrast, just above the threshold,
H-NS dimers form 3-dimensional clusters, which are able to connect DNA sites that are
distant from the genomic point of view (DNA-bridging mode). The model provides clear
rationales for the experimental observations that the switch between the two modes is a
threshold effect and that the ability of H-NS dimers to form higher order oligomers is crucial
for their bridging capabilities.
(#) [email protected]
1
INTRODUCTION
Nucleoid Associated Proteins (NAP) are DNA-binding proteins with low to medium
sequence specificity, which participate actively in bacteria' chromosome organization and
gene expression regulation by bridging, bending, or caging the DNA molecule (1-3). Despite
continuous efforts of several groups (see (1-3) and references therein), the mechanisms by
which NAP achieve these goals remain in most cases elusive and controversial. Part of the
difficulty arises from the fact that these mechanisms appear to be specific to each NAP. For
example, the histone-like nucleoid structuring protein (H-NS), ParB, and structural
maintenance of chromosome proteins (SMC) are all believed to bridge DNA, but the bridges
are qualitatively different, which results in different effects on chromosome organization and
gene expression regulation (4). Even worse, a single NAP may display several binding modes
to DNA, depending on a variety of external factors, like pH, temperature, and the composition
of the buffer (5-9). For example, atomic force microscopy and optical tweezers experiments
first suggested that H-NS binding to DNA leads to the formation of bridges between DNA
duplexes (this is the so-called bridging mode) (5), while subsequent magnetic tweezers
experiments instead concluded that the DNA molecule rather adopts a more extended and
stiffer conformation upon binding of the proteins (this is the so-called stiffening mode) (6). It
was later pointed out that the discrepancy between the two sets of experiments may arise from
the fact that the buffer used in the first set of experiments contained divalent salt cations,
while the buffer used in the second set did not, and it was accordingly shown that both
binding modes do exist and that a switch between the two of them can be driven by changes
in divalent cations concentrations (7). More recent small angle neutron scattering experiments
(8), as well as experiments performed in confined geometries (9), confirmed the crucial role
of divalent cations. Still, it should be pointed out, as stated in (7), that "the specific
mechanism by which magnesium and calcium ions alter H-NS binding properties is currently
unknown", which is an all the more regrettable lack, as the regulation properties of H-NS
(10,11), as well as other NAP (12,13), are probably linked to their DNA-binding properties.
The purpose of the present paper is to propose an explanation for the role of divalent cations,
based on our current knowledge of the properties of H-NS proteins and the results of
simulations performed with a coarse-grained model, which was developed specifically for this
purpose.
H-NS is a small protein (137 residues, 15.5 kDa), which is functional as a dimer. Each
monomer consists of a N-terminal oligomerization domain (residues 1-64) (14,15) and a C-
2
terminal DNA-binding domain (residues 91-137) (16) connected by a flexible linker (17). A
secondary dimerization site at the C-terminal end of the main oligomerization domain allows
H-NS dimers to organize in superhelical chains in crystals (18) and is probably also
responsible for the oligomerization of H-NS in solution, where dimers, tetramers, and larger
oligomers have been observed under different conditions (14,19,20). A thermodynamic
analysis of such solutions led to a value of the enthalpy change upon dimerization or
tetramerization of H-NS proteins of the order of
10
TkB
at room temperature (20), which is
very close to the value that was reported for the enthalpy change upon formation of a complex
between DNA and a H-NS protein in solution (21). As will be developed below, this
similarity of the values of the enthalpy changes upon oligomerization of H-NS and binding of
H-NS to DNA is crucial for the dynamics of H-NS/DNA mixtures.
A second important point deals with the very peculiar properties of DNA when
immersed in a solution containing dilute cations. Naked DNA is indeed a highly charged
polyanion with two phosphate groups per base pair, resulting in a bare linear charge density of
about -5.9
nm/e
, where e is the absolute charge of the electron. However, the Manning-
Oosawa condensation theory (22,23) stipulates that the net linear charge density along a
polyion immersed in a buffer containing counterions of valency Z cannot be larger than
BZe
ℓ
/(
)
, where
=ℓ
B
2
e
pe
4/(
e
0
Tk
Br
)
denotes the Bjerrum length of the buffer, that is, the
distance at which the electrostatic interaction between two elementary charges e is equal to
the thermal energy TkB
. If the bare charge density along the naked polyion is larger than the
critical value
BZe
ℓ
/(
)
, then an instability occurs and counterions coalesce on the polyion and
neutralize an increasing number of its charges, till the net density reduces to the critical value
BZe
ℓ
/(
)
. Owing to its relative dielectric constant
e
r
80
, the Bjerrum length of water at
25°C is
ℓ
B
7.0»
nm and the critical charge density
nm/4.1 e
(for a buffer with
monovalent cations) or
nm/7.0 e
(for a buffer with divalent cations). This implies that (i)
counterions do coalesce on the DNA molecule for both monovalent and divalent cations, and
(ii) increasing the valency of the cations leads to a proportional reduction of the net charge
density along the DNA.
Counterion condensation is expected to have important consequences on H-NS/DNA
interactions. Indeed, except in the vicinity of a few high-affinity binding sites (24), H-
NS/DNA interactions are mainly non-specific, as indicated by the very small variation of the
change in heat capacity with temperature in the range 10-25°C (21) and by the fact that
protein occupation on DNA decreases as monovalent salt concentration increases (7). H-NS
3
»
-
»
-
»
proteins and cationic counterions therefore compete for binding to the DNA and theory
indicates that the binding of polypeptides to the DNA is indeed accompanied by the release of
counterions in the buffer (25-28). The counterions regain translational entropy upon release
from the DNA (25-28), so that the net energy balance for the binding of ligands to the DNA
results from subtle enthalpy-entropy compensations (29,30). Of outmost importance for the
present work is the fact that this balance depends sensitively on the valency of the cationic
counterions present in the buffer. It has indeed repeatedly been observed that addition of small
amounts of divalent cations provokes a substantial decrease of the free energy of binding of
polypeptides and proteins to nucleic acids (31-33).
In contrast, the bare charge density is much smaller along proteins than along nucleic
acids, usually smaller
than Manning-Oosawa's critical density, so
that counterion
condensation is expected to be much less marked for proteins than for nucleic acids, although
there are theoretical indications that it might not totally vanish (34). Still, the short length of
H-NS proteins should contribute to further diminish the eventual importance of counterion
condensation (34), if it does take place. As a consequence, H-NS/H-NS interactions are
expected to remain mostly insensitive to the valency of the cations of the buffer. More
precisely, both H-NS/H-NS and H-NS/DNA interactions depend on the ionic strength of the
buffer through the variation of the Debye length, but H-NS/DNA interactions depend
additionally very sensitively on the valency of the cations through the mechanisms of
counterion condensation and release.
Considered together, the arguments sketched above consequently suggest that binding
of H-NS to DNA is favored in buffers with monovalent cations, while oligomerization of H-
NS proteins is favored in buffers containing substantial amounts of divalent cations, the role
of divalent cations (like magnesium and calcium ions) consisting precisely in displacing the
equilibrium from predominant H-NS/DNA interactions towards predominant H-NS/H-NS
interactions.
While this assumption is the key point of the present work, there still remains to
understand why displacing the equilibrium towards leading H-NS/H-NS interactions drives
H-NS from the stiffening to the bridging mode. In the same spirit as for previous work dealing
with facilitated diffusion (35-37) and the compaction of genomic DNA inside the nucleoid
(38-42), a mesoscopic beads and springs-type model was developed specifically to answer
this question. This model incorporates the effective valency of the cations, Z, as a free
parameter and simulations were launched with different values of Z. The results presented in
Results and Discussion highlight the fact that a small variation of Z is indeed able to induce
4
strong changes in the conformations of the system. More precisely, in the range of values of Z
where H-NS/DNA interactions prevail over H-NS/H-NS interactions, the dynamics of the
system is driven by the attachment of protein chains to the DNA chain in cis configuration,
their sliding along the DNA, and the formation of filaments of proteins along the DNA
duplex, with the length and internal connectivity of the filaments increasing with Z. Such
filaments have repeatedly been reported as a characteristic of the stiffening mode (6,7,10). For
slightly larger values of Z, H-NS oligomerization however takes over the binding of H-NS to
DNA and the system rather organizes in the form of 3-dimensional H-NS clusters, which
efficiently bind regions of the DNA duplex that are broadly separated from the genomic point
of view. This is the bridging mode, which leads to the collapse of the DNA for sufficiently
large protein concentrations.
This work consequently supports a mechanism for the switch of H-NS proteins from
the DNA-stiffening mode to the DNA-bridging mode upon increase of the concentration of
divalent cations, which consists of (i) the displacement of the equilibrium from predominant
H-NS/DNA interactions towards predominant H-NS/H-NS interactions, and (ii) the resulting
switch of the preferred organization of H-NS proteins from filaments stretching along a
genomic contiguous part of the DNA molecule to clusters able to connect parts of the DNA
molecule that are genomically broadly separated.
METHODS
The mesoscopic model which has been developed for the present study is described in
detail in Model and Simulations in the Supporting Material. In brief, the DNA is modelled as
a circular chain of 2880 beads with radius 1.0 nm separated at equilibrium by a distance 5.2
nm and enclosed in a sphere with radius 120 nm. Two beads represent 15 DNA base pairs.
Both the contour length of the DNA molecule and the cell volume are reduced by a factor of
approximately 200 with respect to their actual values in E. coli cells, so that the nucleic acid
concentration of the model is close to the in vivo one. DNA beads interact through stretching,
bending, and electrostatic terms. The bending rigidity constant is chosen so that the model
reproduces the known persistence length of DNA (50 nm). Electrostatic repulsion between
DNA beads is written as a sum of Debye-Hückel terms, which depend on effective
electrostatic charges placed at the center of each bead. These charges are assumed to be
inversely proportional to a parameter Z, which represents the effective valency of the cations
of the buffer, in order to account for counterion condensation along the DNA molecule. Z is
5
equal to 1 (respectively, 2) for monovalent (respectively, divalent) cations, but may take any
real value between 1 and 2 when divalent cations are added to a buffer that contains
monovalent cations. This point is discussed in more detail in Results and Discussion.
Repulsion between DNA beads decreases as
2
/1 Z upon increase of Z.
H-NS dimers are modeled as chains of 4 beads with radius 1.0 nm separated at
equilibrium by a distance 4.0 nm. For most simulations, 200 protein chains were introduced in
the confining sphere together with the DNA chain, which corresponds to a protein
concentration approximately twice the concentration of H-NS dimers during the cell growth
phase and six times the concentration during the stationary phase (43). Protein chains have
internal stretching and bending energy and interact with each other and with the DNA chain.
The value of the bending constant is assumed to be as low as
TkB2
, in order to account for
the flexible linker that connects the C-terminal and N-terminal domains of H-NS (17).
A major approximation of the model consists in assuming that the interactions among
protein beads and between protein beads and DNA beads are mediated uniquely by effective
electrostatic charges placed at the center of each protein bead and to disregard all interactions
beyond the crude electrostatic ones. The charges are positive for the two terminal beads of
each chain and negative for the two central beads. The terminal beads of each protein chain
can therefore bind either to the beads of the DNA chain or to the central beads of other protein
chains, so that the model accounts for both H-NS oligomerization and binding of H-NS to
DNA. The values of the charges at the center of protein beads are assumed to be independent
of Z, thereby reflecting the fact that counterion condensation on proteins is negligible.
Consequently, the attraction term between terminal beads of a protein chain and central beads
of another protein chain does not depend on Z, while the attraction term between DNA beads
and terminal protein beads evolves as Z/1
. This is one of the key points of the model.
A Lennard-Jones-type excluded volume term is added to the attractive Debye-Hückel
term for pairs of beads with opposite charges. For the sake of simplicity, this excluded volume
potential is assumed to be independent of Z and identical for protein/protein and protein/DNA
pairs of beads. The two parameters of the potential were adjusted manually in order that the
enthalpy changes upon forming a complex between two protein chains and between a protein
chain and the DNA chain are comparable to the experimentally determined value for H-NS
(20,21). As shown in Figs. S1 and S2 of the Supplemental Material, this enthalpy change is
equal to
TkB0.12
both for two protein chains approaching one another perpendicularly, and
for a protein chain approaching the linear DNA chain perpendicularly at
37.1=Z
. Since H-
6
-
NS/DNA attraction decreases like Z/1
, while H-NS/H-NS attraction does not depend on Z, it
is expected that this value
Z =
1.37
plays a critical role in the model, with the formation of H-
NS/DNA bonds being energetically favored for
favored for
Z >
1.37
. This particular value
Z <
37.1=Z
1.37
and oligomerization of H-NS being
will therefore be labelled
critZ in the
remainder of the paper. The evolution with Z of the enthalpy change upon formation of a H-
NS/DNA bond is shown as a solid line in Fig. 1 and the enthalpy change upon formation of a
H-NS/H-NS bond as a horizontal dot-dashed line. The dashed line indicates the energy of the
saddles that separate two minima, see Fig. S2. The distance between the solid and dashed
lines therefore represents the gap that proteins must overcome to translocate from one DNA
binding site to the next one.
It is worth emphasizing that the model proposed here relies on a rather crude
approximation, in the sense that it ignores most of the complexity of the binding of H-NS
proteins to the DNA. The model indeed assumes that binding dynamics is governed by
electrostatic interactions between protein beads and the DNA chain, which carries the critical
charge density for a given value of Z, and disregards more complex mechanisms, like the
release of counterions. This model was used in spite of its naivety, because it nonetheless
reproduces the marked decrease of the H-NS/DNA binding energy upon addition of
multivalent cations, which is central to this work.
The dynamics of the complete system was investigated by integrating numerically the
Langevin equations of motion with kinetic energy terms dropped and with time step set to 1.0
ps for simulations with 200 protein chains and 0.5 ps for 1000 protein chains. After each
integration step, the position of the centre of the confining sphere was slightly adjusted so as
to coincide with the centre of mass of the DNA molecule. Temperature T was assumed to be
33.1=Z
298 K throughout the study. Representative snapshots for 200 protein chains and
,
1.42, and 1.50, are shown in Fig. 2.
RESULTS AND DISCUSSION
As mentioned in the Introduction, the Manning-Oosawa condensation theory (22,23)
stipulates that the effective linear charge along a highly charged polyion immersed in a buffer
containing counterions of valency Z is
BZe
ℓ
/(
)
as soon as counterion condensation occurs.
The case of a buffer containing cations of two different valencies
2Z is more complex,
because the cations compete for condensation. Still, the fractions of DNA phosphate sites
1Z and
7
neutralized by cations of type 1 and 2 ( 1
the equations derived in (44). One can then compute the effective valency of the cations of the
q , respectively) can be readily obtained from
q and
2
buffer, according to
=
Z
ℓ
B
Z
11
b
(
q
1
17.0=b
q
Z
22
where
+
q
Z
11
,
)
q
22
Z
(1)
nm is the effective distance between two charges along naked DNA, and
the total fraction of DNA phosphate charges neutralized by the different cations. Z
is always comprised between
1Z and
2Z and the net linear charge density along the DNA
backbone is
BZe
ℓ
/(
)
. Application of Eq. (1) to a buffer containing 10mM of MgCl2 in
addition to 60mM of KCl, as was used in (5) and also investigated in (31-33), leads to an
effective valency
63.1=Z
, meaning that addition of a relatively small amount of divalent
cations results in a significant increase of Z. Addition of this amount of MgCl2 also increases
the Debye length by about 30%, but this variation affects H-NS/DNA and H-NS/H-NS
interactions essentially in the same way, so that this effect will be disregarded in the present
simulations, which focus on the effects of varying Z on the dynamics of the system.
In the model, the effective valency Z is actually considered as a free parameter, which
affects only the charge placed at the center of each DNA bead, this charge being obtained as
the product of the equilibrium distance between two beads and the net linear charge density
BZe
ℓ
/(
)
. Simulations were run for 9 different values of Z ranging from 1.0 to 1.67 and two
different numbers of protein chains (200 and 1000), while keeping all other parameters of the
system constant. Representative snapshots of the conformations obtained with 200 proteins
chains and Z=1.33, 1.42, and 1.50, are shown in Fig. 2. It is reminded that H-NS/DNA
, while H-NS/H-NS interactions are
interactions are energetically favored for
< ZZ
37.1
=
crit
favored for
Z >
critZ
. Visually, one indeed observes a qualitative difference between the three
snapshots in Fig. 2, with the protein chains organizing principally in short filaments stretching
along the DNA for Z=1.33, in rather loose clusters connected to the DNA for Z=1.42, and in
bigger, more compact clusters connected to the DNA for Z=1.50.
A more quantitative insight into the evolution of the conformations of the system can
be gained by plotting, for increasing values of Z, the probability distribution
)(sp
for a
protein chain to be bound to s DNA beads (Fig. 3, left column), the probability distribution
( )q s for a protein chain to be bound to s other protein chains (Fig. 3, right column), as well as
the probability distribution
( )u s for this protein chain to belong to a cluster composed of s
8
-
-
-
-
protein chains (Fig. 4) (the closely related plot of the size distribution of protein clusters is
shown is Fig. S3 of the Supporting Material). For this purpose, it was considered that the
protein chain is bound to a DNA bead if their interaction is attractive and of total magnitude
(computed according to Eq. (S11)) larger than
B3 k T , and that two protein chains are bound if
the interaction between one chain and one terminal bead of the other chain is attractive and of
total magnitude (computed according to Eq. (S9)) larger than
B3 k T . The choice of the
threshold is somewhat arbitrary but the principal features of the distributions do not depend
critically thereon. It is seen in Fig. 3 that for Z=1.00, that is, when H-NS/DNA interactions are
energetically favored with respect to H-NS/H-NS ones, most protein chains bind to four DNA
beads but do not bind to other protein chains. Remembering that tight binding of a protein
chain to the DNA chain involves two consecutive DNA beads (see Fig. S2), this indicates that
most protein chains bind the DNA chain with their two terminal beads (that is, in cis) while
remaining separated from the other protein chains. This is confirmed by the distribution of
cluster sizes (Fig. 4), which indicates that about 70% of the protein chains belong to clusters
of size 1 (and about 25% to clusters of size 2). In contrast, for Z=1.67, that is, when H-NS/H-
NS interactions are energetically favored with respect to H-NS/DNA ones, most protein
chains do not bind to the DNA chain but bind instead on average to 5 other protein chains
(Fig. 3). This suggests that protein chains form large clusters, which interact only loosely with
the DNA chain. The distribution of cluster sizes (Fig. 4) accordingly indicates that large
clusters with size up to 100 chains sequestrate the majority of protein chains.
For 200 protein chains, the transition between these two limiting regimes occurs
around
Z
crit
=
37.1
, as may be checked in Fig. 3. In this range of values of Z, many protein
chains bind to one other protein chain and form simultaneously one strong bond (involving
two DNA beads) with the DNA chain. This suggests that filaments of connected protein
chains form along the DNA chain, where each protein chain binds with one terminal bead to
the DNA chain and with the other terminal bead to the neighboring protein chain. The plot of
( )u s in Fig. 4 accordingly indicates that, for Z=1.33, about 35% of the protein chains belong
to filaments of size 3 to 6 (and about 35% to filaments of size 2). These filaments are
admittedly not very long and they are furthermore dispersed rather randomly over the whole
contour length of the DNA chain. It should however be reminded that 200 protein chains
correspond approximately to one H-NS dimer per 100 DNA base pair, that is, about twice the
physiological concentration during the cell growth phase (43). This concentration is much
smaller than the typical protein concentrations used in (6,7,9,10), which were of the order of
9
one H-NS dimer per one to ten DNA base pairs and led to the collapse of the DNA molecule
when no stretching force was applied to its extremities. Simulations indicate that, for 1000
protein chains, filaments form at lower values of Z, are much longer, and cover significantly
larger portions of the DNA contour length, as may be checked in Fig. 5, which shows a
representative snapshot of the system for Z=1.17.
Careful analysis of the results of simulations indicate that the mechanical properties of
the DNA chain (like its persistence length) are not significantly altered by the assembly of
protein filaments on the DNA chain, even for 1000 protein chains. This is probably due to the
fact that the bending rigidity of the protein chains was assumed to be much smaller than the
rigidity of the DNA chain, in order to account for the flexible linker that connects the C-
terminal and N-terminal domains of H-NS (17). As a consequence, the protein chains and the
filaments merely adapt to the deformations of the DNA chain without hampering them
significantly. This property of the model is in clear contradiction with experimental results,
which indicate that the formation of filaments is accompanied by an increase in the
persistence length and stiffness of the DNA molecule (6,7,10). Such a discrepancy between
the results of simulations and experiments strongly suggests that protein filaments are in
themselves not responsible for the increased stiffness and that the mechanism behind it is
more probably related to the way H-NS dimers bind to the DNA duplex. More precisely, it
has been shown that H-NS dimers insert one C-terminal loop inside the minor groove of
double-stranded DNA (45). Binding of a H-NS dimer to a DNA site consequently decreases
the flexibility of the DNA chain at this particular location, a point not accounted for in the
model, and the many H-NS/DNA bonds associated with the formation of protein filaments are
probably responsible for the observed increase in overall DNA stiffness.
Another important feature of protein filaments is that they stretch along the DNA
chain but only seldom bridge sites that are broadly separated from the genomic point of view.
However, the geometry of protein clusters changes drastically above
Z
=
crit
37.1
, where H-
NS/H-NS interactions become energetically favored with respect to H-NS/DNA ones. Indeed,
above
critZ protein clusters grow in three dimensions in the buffer, instead of growing in one
dimension along the DNA chain, which enables them to form simultaneous contacts with
several sites along the DNA chain that are not contiguous from the genomic point of view, as
may be checked in the middle and bottom vignettes of Fig. 2. Particularly striking is the plot,
as a function of Z, of the number of indirect connections between pairs of DNA beads
mediated by H-NS clusters, which is shown in Fig. 6. Two DNA beads are considered to be
10
connected if they are linked by protein chains that form a continuous series of contacts of the
form d-p-p-p-... -p-p-p-d, where d denotes a DNA bead and p a protein chain, and there is no
limit on the number of intercalated protein chains p. The same threshold as above is used for
defining a contact (
B3 k T for both DNA/H-NS and H-NS/H-NS contacts), and only pairs of
DNA beads d separated along the DNA chain (genomic separation) by 50 or more other DNA
beads are taken into account (the result is essentially independent of the exact value of the
separation threshold, as soon as it is larger than the contour length of the short protein
filaments observed for 200 protein chains). It is seen in Fig. 6, that the number of connections
increases sharply just above
critZ , which indicates that connection of distant DNA sites by H-
NS proteins is efficient only when H-NS/H-NS interactions are favored with respect to H-
NS/DNA ones and H-NS proteins are able to form clusters in three dimensions.
Let us mention for the sake of completeness that other quantities, like the coverage of
the DNA chain by protein chains, may however display smoother variations around
critZ than
the quantities discussed above, as can be checked in Fig. S4 of the Supporting Material. For
the specific case of the coverage of the DNA chain by protein chains, the smoother behavior
is due to the fact that the formation of both protein filaments and 3-dimensional clusters
contribute to reduce the coverage of DNA (because of the overlap of protein chains in the
former case and of protein chains not being bound to the DNA chain in the latter case), so that
coverage of DNA decreases steadily and rather uniformly over the full range of variation of Z.
It is worth emphasizing that the present work consequently describes the DNA-
bridging mode of H-NS proteins as consisting of protein clusters that bind broadly separated
sites along the DNA chain, and not, as is often more or less implicitly assumed, as tracts of
parallel proteins that bridge two parallel DNA duplexes. This latter description emerges quite
naturally when looking at atomic force micrographs of DNA/protein complexes deposited on
mica plates (5), but simulations have shown that the conformations of DNA/protein
complexes are significantly different in bulk and in planar conditions and that the parallel
protein bridges geometry arises from the rearrangement of the complexes after their
deposition on the charged surface (38,39). Moreover, recent single-molecule experiments
have reported the presence of clusters of H-NS proteins buried inside the nucleoid (46), while
other experiments have shown that mutants of H-NS, which have disrupted dimer-dimer
interactions and cannot form higher order oligomers, are also unable to form bridges between
two DNA duplexes (47). The ability for H-NS dimers to clusterize therefore appears as a
prerequisite for their ability to bridge distant DNA sites, which strongly supports the
11
conclusions of the present work. Finally, it has been shown in the same work that the
transition from no bridging to complete bridging is very abrupt upon increase of the
concentration of magnesium ions (47), which is again in agreement with the threshold effect
observed in Fig. 6.
CONCLUSION
The coarse-grained model developed in the present work fully supports the mechanism
proposed in the Introduction, according to which the primary role of divalent cations in the
switch of H-NS proteins from the DNA-stiffening mode to the DNA-bridging mode consists
in decreasing the strength of H-NS/DNA interactions with respect to the strength of H-NS/H-
NS interactions, with the latter ones becoming energetically favored above a certain threshold
of the effective valency. The geometry of H-NS clusters changes drastically at this threshold.
Below the threshold, H-NS dimers form filaments, which stretch along the DNA molecule but
are quite inefficient in bridging DNA sites that are not contiguous from the genomic point of
view. This is the DNA-stiffening mode. In contrast, above the threshold, H-NS dimers form
3-dimensional clusters, which are able to connect DNA sites that are distant from the genomic
point of view. This is the DNA-bridging mode.
The model does not reproduce the increase in DNA stiffness induced by the assembly
of protein filaments on the DNA molecule, because it does not account for the local increase
of rigidity associated with the insertion of the C-terminal loop of H-NS proteins inside the
minor groove of the DNA molecule. In contrast, it provides clear rationales for the
experimental observations that the switch between the two modes is a threshold effect and
that the ability of H-NS dimers to further oligomerize is crucial for their bridging capabilities.
This work suggests that the switch of H-NS proteins between the DNA-stiffening and
DNA-binding modes can be explained on the basis of simple, general arguments. According
to the recent study quoted above (47), it may instead result from a more involved mechanism,
where divalent cations let H-NS dimers switch between a "closed" and an "open" geometry,
which have different oligomerization and DNA-binding properties (47). While it is not clear,
why such a mechanism would display a threshold, it can of course not be excluded that both
mechanisms cooperate to let H-NS proteins switch between the two DNA-binding modes.
Last but not least, it appears that the concentration of divalent cations where the switch
occurs falls within the physiological range, which may be considered as an indication that the
12
mechanism described here may play an important role in the global regulation scheme of
bacterial cells.
SUPPORTING MATERIAL
Model and Simulations section. Figures S1 to S4.
SUPPORTING CITATIONS
Reference (48) appears in the Supporting Material.
13
REFERENCES
1.
Luijsterburg, M. S., M. F. White, R. van Driel, and R. T. Dame. 2008. The major
architects of chromatin: Architectural proteins in bacteria, archaea and eukaryotes. Crit. Rev.
Biochem. Mol. Biol. 43:393-418.
2
Johnson, R. C., L. M. Johnson, J. W. Schmidt, and J. F. Gardner. 2005. Major
nucleoid proteins in the structure and function of the Escherichia coli chromosome. In The
bacterial Chromosome. N. P. Higgins, editor. American Society for Microbiology,
Washington DC. 65-132.
3
Dame, R. T. 2005. The role of nucleoid-associated proteins in the organization and
compaction of bacterial chromatin. Mol. Microbiol. 56:858-870.
4.
Song, D., and J. J. Loparo. 2015. Building bridges within the bacterial chromosome.
Trends in Genetics 31:164-173.
5.
Dame, R. T., C. Wyman, and N. Goosen. 2000. H-NS mediated compaction of DNA
visualised by atomic force microscopy. Nucleic Acids Res. 28:3504-3510.
6.
Amit, R., A. B. Oppenheim, and J. Stavans. 2003. Increased bending rigidity of single
DNA molecules by H-NS, a temperature and osmolarity sensor. Biophys. J. 84:2467-2473.
7.
Liu Y. J., H. Chen, L. J. Kenney, and J. Yan. 2010. A divalent switch drives H-
NS/DNA-binding conformations between stiffening and bridging modes. Gene Dev. 24:339-
344.
8.
Zhang, C., D. Guttula, F. Liu, P. P. Malar, S. Y. Ng, L. Dai, P. S. Doyle, J. A. van
Kan, and J. R. C. van der Maarel. 2013. Effect of H-NS on the elongation and compaction of
single DNA molecules in a nanospace. Soft Matter 9:9593-9601.
9.
van der Maarel, J. R. C., D. Guttula, V. Arluison, S. U. Egelhaaf, I. Grillo, and V. T.
Forsyth. 2016. Structure of the H-NS-DNA nucleoprotein complex. Soft Matter 12:3636-
3642.
10.
Lim, C. J. , S. Y. Lee, L. J. Kenney, and J. Yan. 2012. Nucleoprotein filament
formation is the structural basis for bacterial protein H-NS gene silencing. Sci. Rep. 2:509.
14
11.
Kotlajich, M. V., D. R. Hron, B. A. Boudreau, Z. Sun, Y. L. Lyubchenko, and R.
Landick. 2015. Bridged filaments of histone-like nucleoid structuring protein pause RNA
polymerase and aid termination in bacteria. eLife 4:e04970.
12.
Lim, C. J., Y. R. Whang, L. J. Kenney, and J. Yan. 2012. Gene silencing H-NS
paralogue StpA forms a rigid protein filament along DNA that blocks DNA accessibility.
Nucleic Acids Res. 40:3316-3328.
13.
Lim, C. J., S. Y. Lee, J. Teramoto, A. Ishihama, and J. Yan. 2013. The nucleoid-
associated protein Dan organizes chromosomal DNA through rigid nucleoprotein filament
formation in E. coli during anoxia. Nucleic Acids Res. 41:746-753.
14.
Esposito, D., A. Petrovic, R. Harris, S. Ono, J. F. Eccleston, A. Mbabaali, I. Haq, C. F.
Higgins, J. C. D. Hinton, P. C. Driscoll, and J. E. Ladbury. 2002. H-NS oligomerization
domain structure reveals the mechanism for high-order self-association of the intact protein. J.
Mol. Biol. 324:841-850.
15.
Bloch, V., Y. Yang, E. Margeat, A. Chavanieu, M. T. Augé, B. Robert, S. Arold, S.
Rimsky, and M. Kochoyan. 2003. The H-NS dimerization domain defines a new fold
contributing to DNA recognition. Nat. Struct. Biol. 10:212-218.
16.
Gordon, B. R. G., Y. Li, A. Cote, M. T. Weirauch, P. Ding, T. R. Hughes, W. W.
Navarre, B. Xia, and J. Liu. 2011. Structural basis for recognition of AT-rich DNA by
unrelated xenogeneic silencing proteins. P. Natl. Acad. Sci. USA. 108:10690-10695.
17.
Dorman, C. J., J. C. D. Hinton, and A. Free. 1999. Domain organization and
oligomerization among H-NS-like nucleoid-associated proteins in bacteria. Trends Microbiol.
7:124-128.
18.
Arold, S. T., P. G. Leonard, G. N. Parkinson, and J. E. Ladbury. 2010. H-NS forms a
superhelical protein scaffold for DNA condensation. P. Natl. Acad. Sci. USA. 107:15728-
15732.
19.
Smyth, C. P., T. Lundbäck, D. Renzoni, G. Siligardi, R. Beavil, M. Layton, J. M.
Sidebotham, J. C. D. Hinton, P. C. Driscoli, C. F. Higgins, and J. E. Ladbury. 2000.
Oligomerization of the chromatin-structuring protein H-NS. Mol. Microbiol. 36:962-972.
15
20.
Ceschini, S., G. Lupidi, M. Coletta, C. L. Pon, E. Fioretti, and M. Angeletti. 2000.
Multimeric
self-assembly equilibria
Involving
the histone-like protein H-NS: a
thermodynamic study. J. Biol. Chem. 275:729-734.
21.
Ono, S., M. D. Goldberg, T. Olsson, D. Esposito, J. C. D. Hinton, and J. E. Ladbury.
2005. H-NS is a part of a thermally controlled mechanism for bacterial gene regulation.
Biochem. J. 391:203-213.
22. Manning, G. S. 1969. Limiting laws and counterion condensation in polyelectrolyte
solutions. I. Colligative properties. J. Chem. Phys. 51:924-933.
23.
Oosawa, F. 1971. Polyelectrolytes. Marcel Dekker, New York.
24.
Bouffartigues, E., M. Buckle, C. Badaut, A. Travers, and S. Rimsky. 2007. H-NS
cooperative binding to high-affinity sites in a regulatory element results in transcriptional
silencing. Nat. Struct. Mol. Biol. 14:441-448.
25.
Record, M. T., T. M. Lohman, and P. L. de Haseth. 1976. Ion effects on ligand-nucleic
acid interactions. J. Mol. Biol. 107:145-158.
26.
Record, M. T., C. F. Anderson, and T. M. Lohman. 1978. Thermodynamic analysis of
ion effects on the binding and conformational equilibria of proteins and nucleic acids: the
roles of ion association or release, screening, and ion effects on water activity. Q. Rev.
Biophys. 111:103-178.
27. Mascotti, D., and T. M. Lohman. 1990. Thermodynamic extent of counterion release
upon binding oligolysines to single-stranded nucleic acids. P. Natl. Acad. Sci. USA. 87:3142-
3146.
28.
Fenley, M. O., C. Russo, and G. S. Manning. 2011. Theoretical assessment of the
oligolysine model for ionic interactions in protein-DNA complexes. J. Phys. Chem. B
115:9864-9872.
29.
Breslauer, K. J., D. P. Remeta, W.-Y. Chou, R. Ferrante, J. Curry, D. Zaunczkowski,
J. G. Snyder, and L. A. Marky. 1987. Enthalpy-entropy compensations in drug-DNA binding
studies. P. Natl. Acad. Sci. USA. 84:8922-8926.
16
30. Wang, S., A. Kumar, K. Aston, B. Nguyen, J. K. Bashkin, D. W. Boykin, and W. D.
Wilson. 2013. Different thermodynamic signatures for DNA minor groove binding with
changes in salt concentration and temperature. Chem. Commun. 49:8543-8545.
31.
Record, M. T., P. L. de Haseth, and T. M. Lohman. 1977. Interpretation of monovalent
and divalent cation effects on the lac repressor-operator interaction. Biochemistry 16:4791-
4796.
32.
Lohman, T. M., P. L. de Haseth, and M. T. Record. 1980. Pentalysine-
deoxyribonucleic acid interactions: A model for the general effects of ion concentrations on
the interactions of proteins with nucleic acids. Biochemistry 19:3522-3530.
33. Mascotti, D. P., and T. M. Lohman. 1992. Thermodynamics of single-stranded RNA
binding to oligolysines containing tryptophan. Biochemistry 31:8932-8946.
34.
Nyquist, R. M., B.-Y. Ha, and A. J. Liu. 1999. Counterion condensation in solutions of
rigid polyelectrolytes. Macromolecules 32:3481-3487.
35.
Florescu, A.-M., and M. Joyeux. 2009. Description of non-specific DNA-protein
interaction and facilitated diffusion with a dynamical model. J. Chem. Phys. 130:015103.
36.
Florescu, A.-M., and M. Joyeux. 2009. Dynamical model of DNA protein interaction:
effect of protein charge distribution and mechanical properties. J. Chem. Phys. 131:105102.
37.
Florescu, A.-M., and M. Joyeux. 2010. Comparison of kinetic and dynamical models
of DNA-protein interaction and facilitated diffusion. J. Phys. Chem. A 114:9662-9672.
38.
Joyeux, M., and J. Vreede. 2013. A model of H-NS mediated compaction of bacterial
DNA. Biophys. J. 104:1615-1622.
39.
Joyeux, M. 2014. Equilibration of complexes of DNA and H-NS proteins on charged
surfaces: A coarse-grained model point of view. J. Chem. Phys. 141:115102.
40.
Joyeux, M. 2015. Compaction of bacterial genomic DNA: Clarifying the concepts. J.
Phys.: Condens. Matter. 27:383001.
41.
Joyeux, M. 2016. In vivo compaction dynamics of bacterial DNA: A fingerprint of
DNA/RNA demixing ? Curr. Opin. Colloid Interface Sci. 26:17-27.
17
42.
Joyeux, M. 2017. Coarse-grained model of the demixing of DNA and non-binding
globular macromolecules. J. Phys. Chem. B 121:6351-6358.
43.
Azam, T. A., A. Iwata, A. Nishimura, S. Ueda, and A. Ishihama. 1999. Growth phase-
dependent variation in protein composition of the Escherichia coli nucleoid. J. Bacteriol.
181:6361-6370.
44. Wilson, R. W., and V. A. Bloomfield. 1979. Counterion-induced condensation of
deoxyribonucleic acid. A light-scattering study. Biochemistry 18:2192-2196.
45.
Gordon, B. R. G., Y. Li, A. Cote, M. T. Weirauch, P. Ding, T. R. Hughes, W. W.
Navarre, B. Xia, and J. Liu. 2011. Structural basis for recognition of AT-rich DNA by
unrelated xenogeneic silencing proteins. P. Natl. Acad. Sci. USA. 108:10690-10695.
46.
Stracy, M., C. Lesterlin, F. G. de Leon, S. Uphoff, P. Zawadski, and A. N. Kapanidis.
2015. Live-cell superresolution microscopy reveals the organization of RNA polymerase in
the bacterial nucleoid. Natl. Acad. Sci. USA. 112:E4390-E4399.
47.
van der Valk, R. A., J. Vreede, L. Qin, G. F. Moolenaar, A. Hofmann, N. Goosen, and
R. T. Dame. 2017. Mechanism of environmentally driven conformational changes that
modulate H-NS DNA-bridging activity. eLife 6:27369.
48.
Jian, H., A. Vologodskii, and T. Schlick. 1997. A combined wormlike-chain and bead
model for dynamic simulations of long linear DNA. J. Comp. Phys. 136:168-179
18
FIGURE CAPTIONS
Figure 1 : Plot, as a function of the effective valency Z of the cations of the buffer, of the
minimum of the potential energy felt by a linear protein chain approaching the linear DNA
chain perpendicularly (solid red curve) and of the energy of the saddle separating two minima
(dashed blue curve). The potential energy surface itself is shown in Fig. S2 of the Supporting
Material for
Z
= Z
crit
=
37.1
. The dot-dashed green horizontal line represents the minimum of
the potential energy felt by a linear protein chain approaching another linear protein chain
perpendicularly. The potential energy surface itself is shown in Fig. S1 of the Supporting
Material. The solid and dot-dashed lines intersect at
Z
= Z
crit
=
37.1
.
Figure 2 : Representative snapshots of the system with 200 proteins chains for Z=1.33 (top),
1.42 (middle) and 1.50 (bottom). The blue line connects the centers of successive DNA beads.
Red (respectively, green) spheres represent terminal (respectively, central) beads of protein
chains. The confining sphere is not shown.
Figure 3 : Probability distribution
)(sp
for a protein chain to be bound to s DNA beads (left
column) and probability distribution
( )q s for a protein chain to be bound to s other protein
chains (right column), for values of Z increasing from 1.00 to 1.67. Note that binding of a
protein chain to the DNA chain usually involves two successive DNA beads. Each plot was
obtained from a single equilibrated simulation with 200 protein chains by averaging over a
time interval of 0.1 ms.
Figure 4 : Decimal logarithm of
( )u s , the probability distribution for a protein chain to
belong to a cluster containing s protein chains, for values of Z increasing from 1.00 to 1.67.
Each plot was obtained from a single equilibrated simulation with 200 protein chains by
averaging over a time interval of 0.1 ms.
Figure 5 : Representative snapshots of the system with 1000 proteins chains for Z=1.17. The
blue line connects the centers of successive DNA beads. Red (respectively, green) spheres
represent terminal (respectively, central) beads of protein chains. The confining sphere is not
shown.
19
Figure 6 : Plot, as a function of Z, of the average number of connections between pairs of
DNA beads mediated by protein chains, for 200 protein chains. Only pairs of beads separated
along the chain by 50 (or more) other DNA beads are taken into account.
20
Figure 1
21
Figure 2
22
Figure 3
23
Figure 4
24
Figure 5
25
Figure 6
26
Role of salt valency in the switch of H-NS proteins
between DNA-bridging and DNA-stiffening modes
M. Joyeux
LIPhy, CNRS and Université Grenoble Alpes, Grenoble, France
MODEL AND SIMULATIONS
Temperature T is assumed to be 298 K throughout the study. The model for the DNA
molecule consists of a circular chain of
=n
2880
beads with radius
0.1=a
nm separated at
equilibrium by a distance
=l
0
5.2
nm and enclosed in a sphere with radius
=R
0
120
nm.
Two beads represent 15 DNA base pairs. The contour length of the DNA molecule and the
cell volume correspond approximately to 1/200th of the values for E. coli cells, so that the
nucleic acid concentration of the model is close to the physiological one. The potential energy
of the DNA chain consists of 4 terms, namely, the stretching energy
sV , the bending energy
bV , the electrostatic repulsion
eV , and a confinement term wallV
=
E
DNA
+
VV
b
s
+
V
e
+
V
wall
.
The stretching and bending contributions write
n
∑
(
l
k
2
l
0
)
=
V
s
=
V
b
h
2
g
2
k
=
1
n
∑
=
1
k
q
2
k
,
(S1)
(S2)
where
kr denotes the position of DNA bead k,
kl
=
r
k
r
k
1+
the distance between two
successive beads, and
q
k
=
r
arccos((
k
r
k
+
1
r
)(
k
+
1
r
k
+
2
/()
r
k
r
k
+
1
r
k
+
1
r
k
+
2
))
the angle
formed by three successive beads. The stretching energy
sV is a computational device without
biological meaning, which is aimed at avoiding a rigid rod description. The stretching force
constant h is set to
h =
100
/
lTk
B
2
0
, which insures that the variations of the distance between
successive beads remain small enough (1). In contrast, the bending rigidity constant is
27
-
-
-
-
-
-
obtained from the known persistence length of the DNA,
50=x
nm, according to
= x
g
/
lTk
B
0
=
20
Tk
B
.
Electrostatic repulsion between DNA beads that are not close neighbours along the
chain is written as a sum of repulsive Debye-Hückel terms with hard core
=
(
V
e
e
DNA
Z
2
)
n
n
4
∑ ∑
=
1
k
+=
kK
(
r
kH
4
r
K
)
,
where
)(
rH
=
1
pe
4
exp
r
r
a
2
Dr
.
(S3)
(S4)
Interactions between close neighbours (
1
Kk
3
) are not included in Eq. (S3) because it
is considered that they are already accounted for in the stretching and bending terms.
e =
denotes the dielectric constant of the buffer. The value of the Debye length Dr is set
80 e
0
to 3.07 nm, which corresponds to a concentration of monovalent salt of 0.01 M. DNAe
is the
electric charge, which is placed at the centre of each DNA bead when considering that the
buffer contains only monovalent cations (
1=Z
). The numerical value
-=
e
DNA
.3
525
e
,
where e is the absolute charge of the electron, is the product of 0l and the net linear charge
density derived from Manning's counterion condensation theory (
e
/
ℓ
B
41.1
e
nm/
, see
the main text). The charge placed at the center of each bead reduces to
e
DNA
/
Z
when
considering that the buffer contains cations of effective valency
1>Z
.
Finally, the confinement term wallV is taken as a sum of repulsive terms
=
V
wall
10
Tk
B
n
∑
=
1
k
f
(
r
k
)
,
where f is the function defined according to
if
0Rr £
:
)( =rf
0
if
0Rr >
:
)(
rf
=
r
R
0
6
-
1
.
(S5)
(S6)
H-NS dimers are modeled as chains of 4 beads with radius a separated at equilibrium
by a distance
=L
0
0.4
nm. For each protein chain j, charges
e
j
=
=
e
j
1
4
-==
e
j
3
3
e
are placed at the
3
e
at the centre of
centre of terminal beads
central beads
2=m
and
and
4=m
, and charges
e
j
2
1=m
3=m
. The values of these effective charges were obtained from a
naive counting of the number of positively and negatively charged residues in published
28
-
-
-
-
£
-
£
-
»
-
crystallographic structures (2). It is considered that the density of charges along the naked
protein chain is small enough for counterion condensation not to take place in the range
£ Z
the DNA chain or to the central beads of other protein chains, so that the model accounts for
. Moreover, the terminal beads of each protein chain can bind either to the beads of
2
1
both H-NS oligomerization and binding of H-NS to the DNA chain. In most simulations,
=P
which corresponds to a protein concentration approximately twice the concentration of H-NS
protein chains were introduced in the confining sphere together with the DNA chain,
200
dimers during the cell growth phase and six times the concentration during the stationary
phase (3).
The potential energy of the protein chains consists of 4 terms
E
P
=
V
)P(
s
+
)P(
V
b
+
)P/P(
V
e
+
V
(P)
wall
,
(S7)
where the stretching, bending, and confining energies are very similar to their DNA
counterparts
2
jm
(S8)
h
2
G
2
P
3
∑∑
=
1
j
m
=
1
=
P
m
2
∑∑
=
1
j
=
1
m
=
=
=
)P(
V
s
)P(
V
b
V
(P)
wall
(
L
jm
2
L
0
)
10
Tk
B
P
4
∑∑
=
1
j
=
1
m
f
(
R
jm
,)
with
jmR the position of bead m of protein chain j,
jmL the distance between beads m and
m+1 of protein chain j, and
jm
the angle formed by beads m, m+1, and m+2 of protein chain
j. The value of the bending constant is assumed to be as low as
G
B2=
Tk
, in order to account
for the flexible linker that connects the C-terminal and N-terminal domains of H-NS.
The interaction energy between protein chains,
)P/P(
eV
, is taken as the sum of (attractive
or repulsive) Debye-Hückel terms with hard core and (repulsive) excluded volume terms, with
the latter ones contributing only if the corresponding Debye-Hückel term is attractive
)P/P(
V
e
=
P
∑
=
1
j
Hee
j
4
1
j
(
R
j
1
R
j
4
+
)
4
P
1
4
P
∑∑ ∑ ∑
=
1
j
=
1
m
+=
1
j
J
=
1
M
P
+
c
∑ ∑ ∑ ∑
P
F
3,2
=
1
j m
=
4,1
=
=
1
J M
J
j
(
R
jm
R
JM
,)
Hee
jm
JM
(
R
jm
R
JM
)
where F is the function defined according to
r £
if
0r
:
)(
rF
=
r
0
r
2
a
2
a
(
r
0
r
2
a
2
a
+
1)2
29
(S9)
(S10)
£
Q
-
Q
„
-
-
-
-
-
-
-
-
-
if
r > :
0r
)( =rF
0
,
and 0r denotes the threshold distance below which the excluded volume term, taken as the
repulsive part of a 2h order Lennard-Jones-like function with hard core, creates a repulsion
force between oppositely charged beads. The first term in the right-hand side of Eq. (S9)
insures that the two terminal beads of the same chain do not overlap. The numerical values of
the two parameters of the excluded volume potential,
TkB1=c
and
=r
0
5.3
nm, were
adjusted manually in order that the enthalpy change upon forming a complex between two
protein chains is comparable to the experimentally determined value for H-NS. As shown in
Fig. S1, this enthalpy change is equal to
TkB0.12
for two protein chains at equilibrium
approaching one another perpendicularly, which is close to the experimentally determined
value for H-NS (
TkB2.10
(4)).
Finally, the potential energy describing the interactions between the DNA chain and
the protein chains,
E
DNA/P
, is similarly taken as the sum of (attractive or repulsive) Debye-
Hückel terms with hard core and (repulsive) excluded volume terms, with the latter ones
contributing only if the corresponding Debye-Hückel term is attractive
E
DNA/P
=
e
DNA
Z
4
n
P
∑∑∑
=
1
k
=
1
j
=
1
m
He
jm
(
Rr
k
jm
)
+
c
n
P
∑∑ ∑
=
1
k
=
1
j m
=
4,1
F
(
Rr
k
jm
.)
(S11)
It is emphasized that the attraction term between DNA beads and terminal protein beads
scales as
Z/1
in Eq. (S11), while the attraction term between terminal beads of a protein
chain and central beads of another protein chain does not depend on Z in Eq. (S9). This is a
key point of the model, see the main text. The potential energy felt by protein chains at
equilibrium approaching perpendicularly the linear DNA chain at equilibrium is shown in Fig.
S2 for an effective valency of the cations
37.1=Z
. For this particular value of Z, the enthalpy
change upon binding of a protein chain to the DNA chain is
TkB0.12
, which is equal to the
enthalpy change upon binding of a protein chain to another protein chain, and comparable to
experimentally determined values (
TkB0.11
(5)). In contrast, as illustrated in Fig. 1 of the
main text, binding of protein chains to the DNA chain is favored for
37.1<Z
, while binding
to other protein chains is favored for
37.1>Z
. It may also be noted in this figure that the
evolution of the enthalpy change upon binding of H-NS to DNA deviates slightly from an
Z/1
law, which results from the fact that the excluded volume term is assumed to be
independent of Z in Eq. (S11).
30
-
-
»
-
-
-
-
»
The total potential energy of the system,
potE , is the sum of the energies of DNA and
protein chains and DNA/protein interactions
E
pot
=
E
DNA
+
E
P
+
E
.
DNA/P
(S12)
The dynamics of the model was investigated by integrating numerically the Langevin
equations of motion with kinetic energy terms neglected. Practically, the updated position
vector for each bead (whether DNA or protein),
( +n
)1
jr
, is computed from the current position
vector,
)(n
jr
, according to
n
=+
)1
r
(
j
r
)(
n
j
+
tD
t
Tk
B
+
F
)(
n
j
2
tD
t
x
)(
n
,
(S13)
where the translational diffusion coefficient
tD is equal to
(
TkB
ph
6/()
a
)
and
.0=h
00089
Pa s is the viscosity of the buffer at
arising from the potential energy potE ,
=T
)(nx
298
K.
)(n
jF is the vector of inter-particle forces
a vector of random numbers extracted at each step
n from a Gaussian distribution of mean 0 and variance 1, and
which is set to 1.0 ps for
=P
200
protein chains and 0.5 ps for
the integration time step,
t
=P
chains. After each
1000
integration step, the position of the center of the confining sphere was slightly adjusted so as
to coincide with the center of mass of the DNA molecule.
31
D
D
D
SUPPORTING REFERENCES
1.
Jian, H., A. Vologodskii, and T. Schlick. 1997. A combined wormlike-chain and bead
model for dynamic simulations of long linear DNA. J. Comp. Phys. 136:168-179
2.
Arold, S. T., P. G. Leonard, G. N. Parkinson, and J. E. Ladbury. 2010. H-NS forms a
superhelical protein scaffold for DNA condensation. P. Natl. Acad. Sci. USA. 107:15728-
15732.
3.
Azam, T. A., A. Iwata, A. Nishimura, S. Ueda, and A. Ishihama. 1999. Growth phase-
dependent variation in protein composition of the Escherichia coli nucleoid. J. Bacteriol.
181:6361-6370.
4.
Ceschini, S., G. Lupidi, M. Coletta, C. L. Pon, E. Fioretti, and M. Angeletti. 2000.
Multimeric self-assembly equilibria involving the histone-like protein H-NS. A
thermodynamic study. J. Biol. Chem. 275:729-734.
5.
Ono, S., M. D. Goldberg, T. Olsson, D. Esposito, J. C. D. Hinton, and J. E. Ladbury.
2005. H-NS is a part of a thermally controlled mechanism for bacterial gene regulation.
Biochem. J. 391:203-213.
32
Figure S1 : Potential energy felt by a protein chain aligned along the y axis when approaching
another protein chain elongated along the x axis and centered on (0,0). Both chains are at
equilibrium with respect to their stretching and bending degrees of freedom. The black disks
represent two beads of the protein chain elongated along the x axis. In this geometry, the
potential is symmetric with respect to the y axis, in addition to having rotational symmetry
along the x axis. (x,y) denote the coordinates of the center of the terminal bead of the vertical
chain that lies closest to the horizontal chain. The minimum of the potential energy surface
(
) is located on the y axis. Contour lines are separated by
TkB0.12
TkB2
.
33
-
Figure S2 : Potential energy felt by a protein chain aligned along the y axis when approaching
an infinite DNA chain elongated along the x axis, for an effective valency Z=1.37. Both
chains are at equilibrium with respect to their stretching and bending degrees of freedom. The
black disks represent DNA beads. In this geometry the potential has rotational symmetry
along the x axis. (x,y) denote the coordinates of the center of the terminal bead of the vertical
protein chain that lies closest to the horizontal DNA chain. The minima of the potential
) are located at equal distances from two successive DNA beads.
energy surface (
Contour lines are separated by
TkB0.12
TkB2
.
34
-
Figure S3 : Decimal logarithm of
( )v s , the probability distribution for a protein cluster to
contain s protein chains, for values of Z increasing from 1.00 to 1.67. Each plot was obtained
from a single equilibrated simulation with 200 protein chains by averaging over a time
interval of 0.1 ms.
35
Figure S4 : Plot, as a function of Z, of the fraction of the DNA chain covered by protein
chains, for 200 protein chains. A bead k of the DNA chain is considered to be "covered" by a
protein chain if it is bound to a protein chain or surrounded, in the range [
, by two
beads that are bound to protein chains belonging to the same cluster.
]
5
k
+
k
5,
36
-
|
1304.3789 | 1 | 1304 | 2013-04-13T08:23:41 | Detection of atomic spin labels in a lipid bi-layer using a single-spin nanodiamond probe | [
"physics.bio-ph",
"cond-mat.mes-hall"
] | Magnetic field fluctuations arising from fundamental spins are ubiquitous in nanoscale biology, and are a rich source of information about the processes that generate them. However, the ability to detect the few spins involved without averaging over large ensembles has remained elusive. Here we demonstrate the detection of gadolinium spin labels in an artificial cell membrane under ambient conditions using a single-spin nanodiamond sensor. Changes in the spin relaxation time of the sensor located in the lipid bilayer were optically detected and found to be sensitive to near-individual proximal gadolinium atomic labels. The detection of such small numbers of spins in a model biological setting, with projected detection times of one second, opens a new pathway for in-situ nanoscale detection of dynamical processes in biology. | physics.bio-ph | physics | Detection of atomic spin labels in a lipid bi-layer using a single-spin
nanodiamond probe
Stefan Kaufmann1,2*, David A. Simpson1,3*, Liam T. Hall1, Viktor Perunicic1, Philipp Senn4,5,
Steffen Steinert6, Liam P. McGuinness1, Brett C. Johnson7, Takeshi Ohshima7, Frank Caruso4,
Jörg Wrachtrup6, Robert E. Scholten8, Paul Mulvaney2, Lloyd C. L. Hollenberg1,3+
1 Centre for Quantum Computation and Communication Technology, School of Physics,
University of Melbourne, Victoria 3010, Australia.
2 School of Chemistry and Bio21 Institute, University of Melbourne, Victoria 3010, Australia.
3 Centre for Neural Engineering, University of Melbourne, Victoria 3010, Australia.
4 Department of Chemical and Biomolecular Engineering, University of Melbourne, Victoria
3010, Australia.
5 Bionics Institute, Victoria 3065, Australia.
6 3. Physikalisches Institut, Research Center SCOPE, and MPI for Solid State Research,
University of Stuttgart, Pfaffenwaldring 57, 70569 Stuttgart, Germany.
7 Semiconductor Analysis and Radiation Effects Group, Japan Atomic Energy Agency, Takasaki,
Japan.
8 Centre for Coherent X-Ray Science, School of Physics, University of Melbourne, Victoria
*These authors contributed equally to this work.
3010, Australia
+Corresponding author: Lloyd Hollenberg - School of Physics, University of Melbourne,
Victoria, 3010, Australia. Tel: +61 3 8344 4210, Email: [email protected]
Classification: PHYSICAL SCIENCES: Applied Physical Sciences
1
Key words: nanomagnetometry, diamond, nitrogen-vacancy centre, biophysics, lipid bilayers
Abstract
Magnetic field fluctuations arising from fundamental spins are ubiquitous in nanoscale biology,
and are a rich source of information about the processes that generate them. However, the ability
to detect the few spins involved without averaging over large ensembles has remained elusive.
Here we demonstrate the detection of gadolinium spin labels in an artificial cell membrane under
ambient conditions using a single-spin nanodiamond sensor. Changes in the spin relaxation time
of the sensor located in the lipid bilayer were optically detected and found to be sensitive to near-
individual (4 ± 2) proximal gadolinium atomic labels. The detection of such small numbers of
spins in a model biological setting, with projected detection times of one second, opens a new
pathway for in-situ nanoscale detection of dynamical processes in biology.
The development of sensitive and highly localized probes has driven advances in our
understanding of the basic processes of life at increasingly smaller scales (1). In the last decade
there has been a strong drive to expand the range of probes that can be used for studying
biological systems (2-6), with emphasis on the detection of atoms and molecules in nanometer
sized volumes in order to gain access to information that may be hidden in ensemble averaging.
However, at present there are no nanoprobes suitable for directly sensing the weak magnetic
fields arising from small numbers of fundamental spins in nanoscale biology, occurring naturally
(e.g. free-radicals) or introduced (e.g. spin-labels), which can be a rich source of information
about processes at the atomic and molecular level. Magnetic resonance techniques such as
electron spin resonance (ESR) have played an important role in the development of our
2
understanding of membranes, proteins and free radicals (7); however, ESR sensitivity and
resolution are fundamentally limited to mesoscopic ensembles of at least 107 spins with a
sensitivity of ~ 2 × 109 spins/(Hz)1/2 (8). In a typical ESR application small electron spin label
moieties are attached to the system of interest and their environment is investigated through spin
measurements on the labels. Because of the large ensemble required nanoscopic detail at the few
spin level can be lost in the averaging process. Recently, magnetic resonance force microscopy
techniques have demonstrated single spin detection (9-11), but these require cryogenic
temperatures and vacuum. Here we demonstrate a nanoparticle probe – a nitrogen-vacancy spin
in a nanodiamond – which is situated in the target structure itself and acts as a nanoscopic
magnetic field detector under ambient conditions with non-contact optical readout. We employ
this probe to detect near-individual spin labels in an artificial cell membrane at a projected
sensitivity of ~ 5 spins/(Hz)1/2, effectively bridging the gap between traditional ESR ensemble
based techniques and the ultimate goal of few-spin nanoscale detection under biological
conditions.
Results and discussion
The overall set-up of our experiment is shown schematically in Figure 1A. The nanoparticle
probe is the single spin of a nitrogen-vacancy (NV) centre contained in a nanodiamond particle
(Figure 1B). NV-nanodiamonds are attractive nanoprobes for biology as they exhibit low toxicity
and are photostable (12-14). Here we investigate the magnetic field sensing capabilities of the
NV centre in a biological context. Around these NV-nanodiamonds we formed a supported lipid
bilayer (SLB) on a glass substrate. The SLB was created with lipids labeled with Gd3+, a
common MRI contrast agent with spin 7/2 (Figure 1C) which produce characteristic magnetic
3
fluctuations in the lipid environment (Figure 1D) and form our detection target. The ground state
of the NV centre has a zero-field splitting of D = 2.87 GHz between the 0〉 and ±1〉 spin levels
and optical-based spin state readout is possible due to the significantly lower fluorescence of the
±1〉 compared to the 0〉 state (15) (Figure 1E). These properties have led to demonstrations of
DC and AC magnetic field detection using single NV spins (16-18), and wide-field detection
schemes employing ensembles of NV centres (19-22). For biological applications, where atomic
level processes produce magnetic field fluctuations, it has been proposed that changes in the
quantum decoherence of the NV spin could provide a more sensitive detection mechanism (23-
25). Ensemble relaxation based imaging has been demonstrated with sub-cellular resolution and
detection of ~103 spins (26). Towards the goal of using single NV spins as in situ nano-
magnetometer probes in biology, quantum measurements on single NV spins have been carried
out in a living cell (27), and with the recent detection of external spins in well controlled
environments (28-30), the critical milestone is to demonstrate nanoscopic external spin detection
in a biological context. Here we achieve this with near individual spin sensitivity by monitoring
the relaxation time (T1) of a single NV spin probe (Figure 1E, F) embedded in the SLB target
itself. In addition, our results indicate that the NV spin is sensitive to cross-lipid magnetic
fluctuations arising from the small number of Gd labeled lipids in the SLB in the vicinity of the
nanodiamond (Figure 1A, D).
4
Figure 1: Schematic of nanoscopic detection of spin labels in an artificial cell membrane using a single-spin
nanodiamond sensor. A. A supported lipid bilayer (SLB) is formed around a nanodiamond immobilized on a glass
substrate. B. The nanodiamond contains a single nitrogen-vacancy (NV) optical centre which acts as a single spin
sensor by virtue of the magnetic levels in the ground state. C. Gadolinium (Gd) spin labeled lipids are introduced
into the SLB. D. Magnetic field fluctuations arising from Gd spin labels affect the quantum state of the NV spin,
measured through the NV relaxation time, T1. E. The electronic energy structure of the NV centre showing the
fluorescent cycle and optical spin readout of the spin states 0〉 and ±1〉, and the protocol for the T1 measurement. F.
A schematic illustration of the T1 measurement. The relaxation of the NV spin in the target environment is compared
to that in the reference environment. Measurement at a single time point in the evolution allows faster detection.
The SLB was formed on the nanodiamond-glass substrate in tris-buffered saline (TBS) solution
via the vesicle fusion technique (31) (Figure 2A), and characterized by both fluorescence
recovery after photo-bleaching (FRAP) and atomic force microscopy (AFM) (Figure 2B, and
Supplementary Information). AFM investigations showed that the SLB formed mainly around
the nanodiamond particles (mean size = 15.9 ± 9.5 nm). The T1 time of the NV spin was
5
measured by optically polarizing into the 0〉 state and measuring the probability P0(t) of finding
the NV in the initial 0〉 state at a later time t through spin state fluorescence contrast. In a low
background field the ±1〉 spin states of the NV are approximately degenerate so the P0(t) decays
as:
tP
)(
0
=
1)(3/1(
+
exp(
Tt
/
−
1
)
+
exp(
Tt
/2
−
1
))
.
(1)
This form is a consequence of the 3-state nature of the NV spin manifold and the broad spectrum
of the target spin bath (see Supplementary Information for the derivation). Figure 3A (green
dashed curve) shows P0(t) (normalised fluorescence) for a single NV centre (NV1) under the
TBS control conditions. Fitting to Equation (1) we obtain a reference relaxation time T1[TBS] =
117±11 µs. T1[TBS] includes intrinsic nanodiamond sources of relaxation such as N donors
(electron spin bath), and 13C spins (nuclear spin bath), but is dominated by surface electronic
spins fluctuating in the GHz regime. After the formation of a SLB without Gd spin labels (under
the same TBS conditions), the measurement was repeated on the same NV (green data points) to
give T1[SLB] = 139±18 µs, showing no statistically significant change for the unlabeled lipid
case. Removal of the SLB, which included a 2 min oxygen plasma etching step, was followed by
a new TBS reference T1 measurement (blue dashed curve). Upon formation of the SLB with Gd
spin labeled lipids at 10% (w/w) concentration (blue data points), we observed a 61% reduction
in the NV relaxation time to T1[SLB:Gd] = 48±3 µs.
6
Figure 2: Formation and characterisation of a supported lipid layers (SLB) on nanodiamond/glass substrates. A.
SLBs were formed by the vesicle fusion technique and characterized by fluorescence recovery after photobleaching
(FRAP). B. Confocal images of FRAP measurements directly after the bleaching cycle for a SLB+10% (w/w) of the
Gd spin label and 1% (w/w) of a fluorescent label. For characterisation, three different substrate conditions were
investigated. Confocal and atomic force microscopy images show nanodiamond density and size distribution.
To demonstrate the consistency of the effect, this sequence was performed for five NV centres
(NV1-NV5), in distinct nanodiamonds. Figure 3B shows the percentage changes in the T1 times
for all five NV centres relative to their TBS reference values. For the SLB with no spin labels
there is no statistically significant change verifying that the NV T1 relaxation time remains intact
under control conditions. For the SLB labeled with 10% Gd the average reduction in the NV T1
relaxation time from the reference value is 74 ± 6 %, and is remarkably consistent across the set
7
of nanodiamonds. We next investigated the Gd concentration dependence of the change in the T1
relaxation time. In Figure 3C the T1 curves of a single NV centre (NV6) are given for 10% and
1% Gd concentrations, showing the change in T1 decreasing as the number of proximate Gd
spins is reduced. The percentage changes in T1 from the respective TBS reference values as a
function of Gd concentration are given in Figure 3D for another set of four individual NV centres
(NV6-NV9) confirming this trend.
0% Gd-SLB
10% Gd-SLB
0% Gd-SLB fit
10% Gd-SLB fit
TBS (ref)
TBS (ref)
NV1
B
)
%
(
1
T
∆
75
50
25
0
-25
-50
-75
-100
0% Gd-SLB
10% Gd-SLB
400
500
NV1
NV2
NV3
NV4
NV5
A
1
)
u
.
a
(
e
c
n
e
c
s
e
r
o
u
l
F
0.95
0.9
0.85
0.8
0.75
C
1
)
u
.
a
(
e
c
n
e
c
s
e
r
o
u
l
F
0.95
0.9
0.85
0.8
0.75
0
100
200
300
Time (µs)
10% Gd-SLB
1% Gd-SLB
10% Gd-SLB fit
1% Gd-SLB fit
TBS (ref)
TBS (ref)
NV6
1% Gd-SLB
10% Gd-SLB
D
)
%
(
1
T
∆
50
25
0
-25
-50
-75
-100
-125
0
100
200
300
Time (µs)
400
500
NV6
NV7
NV8
NV9
Figure 3: Detection of spin labeled lipids in a supported lipid bilayer (SLB) using the T1 time of single NV spins in
nanodiamonds. A. Relaxation of the spin of a single NV centre (NV1) in a nanodiamond in a SLB without spin
labels (green) and in a SLB with 10% (w/w) Gd spin labels (blue). B. The percentage change in T1 (relative to TBS)
of five single NV spins (NV1-NV5) in distinct nanodiamonds: SLB+0% Gd (green) and SLB+10% Gd (blue). C.
Relaxation of the spin of NV6 for SLB+ 10% Gd (blue) and SLB+1% Gd (orange). D. Concentration dependence of
8
the percentage change in T1 for a set of centres, NV6-NV9. The data in B and C, after fitting to determine the
relaxation times, have been scaled to the same asymptotic value for ease of presentation. All error bars represent the
fitting uncertainties at the 95% confidence interval. Solid curves are the corresponding fits of the data to Equation
(1) which determine T1. Dashed curves correspond to the fitted data of the corresponding reference TBS
measurement prior to each SLB measurement.
In order to understand these results quantitatively we consider the quantum evolution of the NV
spin in the presence of magnetic field fluctuations arising from the various atomic processes
involving the Gd spins in the SLB (see Supplementary Information). The characteristic
timescales of these fluctuations produce specific changes in the measured T1 relaxation time of
the NV spin from the reference value. In terms of external sources of decoherence, the quantum
evolution of the NV spin is determined by the overall spectral distribution function,
efS
(ω
)
,
which is governed by the characteristic environmental magnetic field fluctuation frequency, fe.
The Gd-labeled lipid environment produces magnetic field fluctuations due to cross-lipid Gd-Gd
spin dipole interactions, motional diffusion of individual Gd-labeled lipids, and intrinsic Gd spin
relaxation effects, each with characteristic frequencies
dipf
,
diff
, and
inf
respectively that
contribute to fe. From
efS
(ω
)
we obtain the NV relaxation time T1[Gd] due to Gd-lipid effects:
T
1
]Gd[
=
(
f
2
e
+
(
D
−
2
D
Gd
2
2/))
2
Bf
e
eff
, where
B
eff
=
63
(
γγµπσ
0
NV
Gd
/)
h
128
2
is the
characteristic Gd-NV dipolar magnetic field interaction, h is the NV depth below the
nanodiamond surface, σ is the areal Gd spin density, and DGd is the zero field splitting parameter
of the Gd spin, with the dominant contribution to T1 coming from the 2DGd transition. We note
that the 4DGd and 6DGd transitions are too far detuned from D to have any significant effect on
T1. We obtain
dipf
by integrating the Gd-Gd dipolar auto-correlation function over a planar SLB
distribution to give
f
dip
=
2
63
(
)
πσγµ
Gd
0
2/3
32/
π
. For the diffusion of Gd-lipids in the SLB we
9
obtain
f
diff
=
9
D
l
16/
h
2
where Dl is the Gd-lipid diffusion constant. The intrinsic Gd spin
relaxation is characterized by
in ≈f
GHz
1
(32). For the NV depths expected in these 16 nm
nanodiamonds and Gd concentrations employed here, the effect of Gd-Gd interactions on the NV
relaxation time dominate over Gd diffusion effects. Expanding around the low Gd density limit
we arrive at a theoretical model for the NV relaxation time T1[Gd], due to proximate Gd labeled
lipids:
T
1
]Gd[
=
4
h
w
2/3
(
aw
+
b
)
,
(2)
where w is the %w/w Gd-lipid concentration, and a and b are constants involving the physical
parameters D, DGd and fin (see Supplementary Information). Because the nanodiamond surface
spins comprise a two-dimensional distribution we also obtain T1[TBS] ∝ h4. Hence there is a
cancelation of the depth dependence in the percentage change in relaxation times,
∆T1[SLB:Gd]/T1[TBS], which may explain the uniformity of the effect across the set of NV-
nanodiamonds shown in Figure 3B.
To directly compare our data with Equation (2) we extract T1[Gd] by subtracting the reference
rate (i.e. Γ1[Gd] = Γ1[SLB:Gd] – Γ1[TBS]) and in Figure 4 plot this quantity for all NV1-NV9
against Gd concentration, w. Generally, the data points lie in a NV depth range consistent with
the measured AFM nanodiamond size distribution indicating that we observe T1[Gd] times in
broad agreement with those predicted by Equation (2). The trend to shallower NV depths as we
move from 10% → 1% Gd-lipid concentration is consistent with the etching step in the
processing of the sample between the 10% and 1% measurements which is likely to have
removed several nm of material from the nanodiamonds. However, we note that for low
10
concentrations we also expect some deviation to the scaling in Equation (2) due to the statistics
of low Gd spin numbers.
1E+06
106
)
1E+05
105
s
µ
(
]
1E+04
104
d
G
[
1
1E+03
103
T
,
e
m
1E+02
102
i
t
n
o
1E+01
101
i
t
a
x
a
1E+00
100
l
e
r
V
1E-01
10-1
N
10-2
1E-02
0.5
NV1
NV3
NV5
NV7
NV9
NV2
NV4
NV6
NV8
13nm
h= 18nm
h= 13nm
h= 8nm
h= 5nm
h= 3nm
50
5
Gd lipid-label concentration (w%)
Figure 4: The NV relaxation time due to the presence of Gd spin labeled lipids, T1[Gd], as a function of Gd lipid
concentration (%w/w) for all measured centres, NV1-NV9 (data points). Theoretical curves (dashed) based on
Equation (2) are given for a range of NV depths, h, with a lower bound h = 2nm corresponding to the photostability
limit (33).
The effective number of spins detected, Neff, for a given Gd concentration w is estimated by
comparing the RMS field expectation 〈B2〉1/2 at the NV contributing to the change in relaxation
time integrated over all Gd spins in the membrane, to that of a single Gd spin at distance h, to
obtain
N
eff
2
πσ (see Supplementary Information). Using the NV depth range h ≈ 8 ± 5
h
)8/3(~
nm from the AFM distribution we arrive at a lower bound estimate of the effective number of
spins detected of 4 ± 2 (1% Gd) and 28 ± 24 (10% Gd). Finally, from the data we can determine
projected single time-point detection times assuming the reference TBS value has been well
11
characterised and the measured fluorescence contrast is normalised. For the case of NV7, which
had a relatively fast value of T1[TBS], to detect the fluorescence change at the 95% CL
corresponding to 1% Gd-SLB at a single time point requires a total measurement time of ~1.3s
(see Supplementary Information).
In conclusion, we have demonstrated nanoscopic magnetic detection of spin labels in an artificial
cell membrane using the T1 relaxation time of a single spin NV-nanodiamond probe. Our results
for Gd-labeled lipid concentrations down to 1% w/w correspond to the detection of near-
individual Gd spin labels, with projected single time point detection times of order 1s. The data
are in broad agreement with cross-lipid Gd-Gd spin interactions as the dominant atomic process
detected. These results highlight the potential of the NV-nanodiamond system as a nanoscopic
magnetic probe in biology which circumvents the fundamental problems associated with
ensemble averaging.
Materials and Methods
Nanodiamond preparation
Nanodiamond were purchased from VanMoppes, Switzerland (SYP 0 - 0.03) and irradiated with
high-energy electrons (2 MeV with a fluence of 1×1018 electrons/cm2) and vacuum annealed at
800 oC for 2 hours at the Japan Atomic Energy Agency. The powder was heated to 425 oC in air
for 3 hours, dispersed in water and sonicated with a high-power sonicator (Sonicator 4000,
Qsonica2) for 30 hours. The suspension was centrifuged at 12000 rcf for 120 s and the
supernatant was removed and used as stock suspension. For the adsorption of nanodiamonds,
12
cleaned glass substrates (No. 1, Menzel) were immersed in a 1 mg/ml solution of polyallylamine
hydrochloride (PAH, 70kDa) in 0.5 M NaCl in Milli-Q water for 5 minutes. After rinsing with
Milli-Q water nanodiamonds were adsorbed in the desired concentration in Milli-Q water for 5
minutes. The glass substrates were treated in a 40 W oxygen plasma (25% oxygen in argon, 40
sccm flow rate) for 5 minutes to remove the polyelectrolyte from the substrate. Atomic force
microscopy (AFM) in air was performed with an Asylum MFP-3D in tapping mode with
cantilevers from Olympus (AC160TS, 42 N/m).
Formation of supported lipid bilayers (SLBs)
All chemicals are of analytical grade reagent quality if not otherwise stated. The lipids (Avanti
Polar Lipids
Inc., USA) 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine
(POPC), 1-
palmitoyl-2-oleoyl-sn-glycero-3-phospho-L-serine (sodium salt) (POPS) and 1-palmitoyl-2-12-
[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl-sn-glycero-3-phosphocholine
(NBD-PC)
were dissolved
in
chloroform
and
the
spin
label 1,2-dimyristoyl-sn-glycero-3-
phosphoethanolamine-N-diethylenetriaminepentaacetic
acid
(gadolinium
salt)
(DMPE-
DTPA(Gd)) was dissolved in the mixture chloroform:methanol:water (65:35:8) (v:v). The lipids
were mixed in the desired ratios and were dried under a stream of nitrogen for 1 hour followed
by
re-suspension
in TBS buffer. TBS buffer was prepared by mixing 10 mM
Tris(hydroxymethyl)-aminomethane (Sigma-Aldrich Pty. Ltd., Australia) and 150 mM sodium
chloride (Sigma-Aldrich Pty. Ltd., Australia) in water with the pH adjusted to 7.4 by stepwise
addition of 1 M hydrochloric acid. Milli-Q water (18.2 MΩ and at most 4 ppm (TOC), Millipore,
USA) was used throughout. The buffer was filtered through 0.2 µm filters (PALL Acrodisc
Syringe Filter) before use. Supported lipid bilayers (SLB) were formed with vesicles extruded 31
13
times through two stacked, 100 nm pore-size polycarbonate membranes (Avanti Polar Lipids
Inc., USA) at a concentration of 0.1 mg/ml with an exposure time of 1 hour (overnight exposure
for quantum measurements in liquid cell). The vesicle size was measured by dynamic light
scattering resulting in an average size of 113 ± 3 nm (s.d). Before the formation of SLBs, 3 mM
Ca +2 (CaCl 2 , Chem-Supply Pty Ltd, Australia) was added to the buffer solution. This improved
the adhesion onto the negatively charged, glass surface (34). After the measurements the liquid
cell was cleaned with 2% (v/w) of sodium dodecyl sulfate in water and isopropyl alcohol and
thoroughly rinsed with water. Before quantum measurements the coverslip was treated in a 40 W
oxygen plasma (25% oxygen in argon, 40 sccm flow rate) for 2 minutes.
T1 measurements
Confocal imaging was performed on a home-built microscope using an oil-immersion objective
(100x, Nikon). T1 measurements were performed using a Spincore PulseBlaster card and a Fast
ComTech P7889 Multiple-Event Time Digitizer card. Typical measurement time for a full
evolution of a T1 curve was 1 hour. Spin relaxation curves were fitted using Equation (1) in the
main manuscript.
Acknowledgements
This work was supported by the Australian Research Council under the Centre of Excellence
(CE110001027) scheme, the Centre for Neural Engineering, ERC Project SQUTEC, EU Project
DINAMO, and the Baden-Wuerttemberg Foundation. SK was supported by a Swiss National
Science Foundation fellowship PBEZP3-133244. FC is supported by the ARC Australian
Laureate Fellowship scheme (FL120100030).
14
References
1.
15.
16.
18.
4.
5.
6.
7.
8.
9.
2.
3.
Michalet X, et al. (2005) Quantum Dots for Live Cells, in Vivo Imaging, and
Diagnostics. Science 307(5709):538-544.
Ament I, Prasad J, Henkel A, Schmachtel S, & Sönnichsen C (2012) Single Unlabeled
Protein Detection on Individual Plasmonic Nanoparticles. Nano Lett. 12(2):1092-1095.
Sorgenfrei S, et al. (2011) Label-free single-molecule detection of DNA-hybridization
kinetics with a carbon nanotube field-effect transistor. Nat. Nanotech. 6(2):126-132.
Heller D A, et al. (2009) Multimodal optical sensing and analyte specificity using single-
walled carbon nanotubes. Nat. Nanotech. 4(2):114.
Li J F, et al. (2010) Shell-isolated nanoparticle-enhanced Raman spectroscopy. Nature
464(7287):392-395.
Matsumoto K, Subramanian S, Murugesan R, Mitchell J B, & Krishna M C (2007)
Spatially resolved biologic information from in vivo EPRI, OMRI, and MRI.
Antioxidants & redox signaling 9(8):1125-1142.
Borbat P, Costa-Filho A, Earle K, Moscicki J, & Freed J (2001) Electron spin resonance
in studies of membranes and proteins. Science 291(5502):266-269.
Blank A, Dunnam C R, Borbat P P, & Freed J H (2004) Pulsed three-dimensional
electron spin resonance microscopy. Appl. Phys. Lett. 85(22):5430-5432.
Rugar D, Budakian R, Mamin H, & Chui B (2004) Single spin detection by magnetic
resonance force microscopy. Nature 430(6997):329-332.
10. Mamin H, Poggio M, Degen C, & Rugar D (2007) Nuclear magnetic resonance imaging
with 90-nm resolution. Nat. Nanotech. 2(5):301-306.
Degen C L, Poggio M, Mamin H J, Rettner C T, & Rugar D (2009) Nanoscale magnetic
resonance imaging. Proc. Natl Acad. Sci. USA 106(5):1313-1317.
Fu C C, et al. (2007) Characterization and application of single fluorescent
nanodiamonds as cellular biomarkers. Proc. Natl Acad. Sci. USA 104(3):727.
Neugart F, et al. (2007) Dynamics of diamond nanoparticles in solution and cells. Nano
Lett. 7(12):3588-3591.
Chang Y R, et al. (2008) Mass production and dynamic imaging of fluorescent
nanodiamonds. Nat Nano 3(5):284-288.
Gruber A, et al. (1997) Scanning confocal optical microscopy and magnetic resonance on
single defect centers. Science 276(5321):2012.
Balasubramanian G, et al. (2008) Nanoscale imaging magnetometry with diamond spins
under ambient conditions. Nature 455(7213):648-651.
17. Maze J R, et al. (2008) Nanoscale magnetic sensing with an individual electronic spin in
diamond. Nature 455(7213):644-648.
Balasubramanian G, et al. (2009) Ultralong spin coherence time in isotopically
engineered diamond. Nat. Mater. 8(5):383-387.
19. Maletinsky P, et al. (2012) A robust scanning diamond sensor for nanoscale imaging with
single nitrogen-vacancy centres. Nat. Nanotech. 7(5):320.
Steinert S, et al. (2010) High sensitivity magnetic imaging using an array of spins in
diamond. Rev. Sci. Instrum. 81(4).
12.
11.
13.
14.
20.
15
25.
26.
Pham L M, et al. (2011) Magnetic field imaging with nitrogen-vacancy ensembles. New
J. Phys. 13(4):045021.
Acosta V M, et al. (2010) Broadband magnetometry by infrared-absorption detection of
nitrogen-vacancy ensembles in diamond. Appl. Phys. Lett. 97(17):174104-174103.
Cole J H & Hollenberg L C L (2009) Scanning quantum decoherence microscopy.
Nanotechnology 20(49):10.
Hall L T, Cole J H, Hill C D, & Hollenberg L C L (2009) Sensing of Fluctuating
Nanoscale Magnetic Fields Using Nitrogen-Vacancy Centers in Diamond. Phys. Rev.
Lett. 103(22):220802.
Hall L T, et al. (2010) Monitoring ion-channel function in real time through quantum
decoherence. Proc. Natl Acad. Sci. USA 107(44):18777-18782.
Steinert S, et al. (2012) Magnetic spin imaging under ambient conditions with sub-
cellular resolution. http://www.arXiv:1211.3242 [cond-mat.mes-hall] .
27. McGuinness L P, et al. (2011) Quantum measurement and orientation tracking of
fluorescent nanodiamonds inside living cells. Nat Nano 6(6):358-363.
28. McGuinness L P, et al. (2012) Ambient Nanoscale Sensing with Single Spins Using
Quantum Decoherence. http://www.arXiv:1211.5749 [cond-mat.mes-hall] .
Grotz B, et al. (2011) Sensing external spins with nitrogen-vacancy diamond. New J.
Phys. 13(5):055004.
30. Mamin H J, Sherwood M H, & Rugar D (2012) Detecting external electron spins using
nitrogen-vacancy centers. http://www.arXiv:1209.3748 [cond-mat.mes-hall] .
31. McConnell H, Watts T, Weis R, & Brian A (1986) Supported planar membranes in
studies of cell-cell recognition in the immune system. Biochim Biophys Acta 864(1):95-
106.
Bertini I, et al. (2004) Persistent contrast enhancement by sterically stabilized
paramagnetic liposomes in murine melanoma. Magnetic resonance in medicine
52(3):669-672.
Bradac C, et al. (2010) Observation and control of blinking nitrogen-vacancy centres in
discrete nanodiamonds. Nat Nano 5(5):345-349.
Rossetti F F, Bally M, Michel R, Textor M, & Reviakine I (2005) Interactions between
titanium dioxide and phosphatidyl serine-containing liposomes: formation and patterning
of supported phospholipid bilayers on the surface of a medically relevant material.
Langmuir 21(14):6443-6450.
29.
32.
34.
33.
21.
22.
23.
24.
16
|
1309.3637 | 2 | 1309 | 2013-10-17T15:23:06 | Nanotribology of biopolymer brushes in aqueous solution using dissipative particle dynamics simulations: an application to PEG covered liposomes in theta solvent | [
"physics.bio-ph",
"cond-mat.soft"
] | We undertake the investigation of sheared polymer chains grafted on flat surfaces to model liposomes covered with polyethylene glycol brushes as a case study for the mechanisms of efficient drug delivery in biologically relevant situations, for example, as carriers for topical treatments of illnesses in the human vasculature. For these applications, specific rheological properties are required, such as low viscosity at high shear rate to improve the transport of the liposomes. Therefore non - equilibrium, DPD simulations of polymer brushes of various length and shear rates are performed to obtain the average viscosity and friction coefficient of the system as functions of the shear rate and polymerization degree under theta solvent conditions, and find that the brushes experience shear thinning at large shear rates.The viscosity and the friction coefficient are shown to obey scaling laws at high shear rate in theta solvent, irrespective of the brushes degree of polymerization. These results confirm recent scaling predictions and reproduce very well trends in measurements of the viscosity at high shear of red blood cells in a liposome containing medium. | physics.bio-ph | physics | Nanotribology of biopolymer brushes in aqueous solution using
dissipative particle dynamics simulations: an application to PEG covered
liposomes in theta solvent
A. Gama Goicochea†1, E. Mayoral1, J. Klapp1,2, and C. Pastorino3
1Instituto Nacional de Investigaciones Nucleares, Carretera México – Toluca s/n, La
Marquesa Ocoyoacac, Estado de México 52750, Mexico
2Departamento de Matemáticas, CINVESTAV del IPN, México D. F. 07360, Mexico
3Departamento de Física, Centro Atómico Constituyentes, CNEA – CONICET, Av.
General Paz 1499, Provincia de Buenos Aires 1650, Argentina
Abstract
We undertake the investigation of sheared polymer chains grafted on flat surfaces to model
liposomes covered with polyethylene glycol brushes as a case study for the mechanisms of
efficient drug delivery in biologically relevant situations, for example, as carriers for topical
treatments of illnesses in the human vasculature. For these applications, specific rheological
properties are required, such as low viscosity at high shear rate to improve the transport of
the liposomes. Therefore, extensive non- equilibrium, coarse – grained dissipative particle
dynamics simulations of polymer brushes of various lengths and shear rates are performed
to obtain the average viscosity and the friction coefficient of the system as functions of the
shear rate and polymerization degree under theta – solvent conditions, and find that the
brushes experience considerable shear thinning at large shear rates. The viscosity () and
† Corresponding author. Electronic mail: [email protected]
1
the friction coefficient () are shown to obey the scaling laws 𝜂~𝛾 −0.31 , and 𝜇~𝛾 0.69 at
high shear rate (𝛾 ) in theta solvent, irrespective of the brushes degree of polymerization.
These results confirm recent scaling predictions and reproduce very well trends in
measurements of the viscosity at high ( 𝛾 ) of red blood cells in a liposome containing
medium.
Keywords: viscosity, friction coefficient, dissipative particle dynamics, biopolymer
brushes, drug-delivering liposomes.
2
I. INTRODUCTION
Polymer brushes, formed when polymer molecules grafted to a surface are stretched, are
important because they play a key role in the stability of colloidal dispersions [1, 2], in the
process of enhanced oil recovery [3], in the design and applications of stimuli – responsive
materials [4], and even in the characterization of the mechanical properties of cancerous
human cells [5], among other reasons [6]. They display some general fundamental
properties, and scaling relations have been derived for them under circumstances such as
high grafting density or negligible chain – chain interaction [7-9]. When polymer brushes
are sheared a plethora of phenomena that are intrinsic to the non – equilibrium nature of the
shearing appear. For example, experiments on fluids compressed between plates under the
influence of external shear have shown that the viscosity of the fluid can be substantially
reduced if a polymer brush is grafted to each plate [10]. Biological examples of sheared
polymer brushes occur in articular cartilage surfaces [11] and synovial joints [12], in
glycocalyx filaments that coat the human circulatory system [13], and in glycosylated cell
surfaces and liposomes, which can be used as carriers for drug delivery [14]. Despite all
these studies, the experimental understanding of the molecular mechanisms that take place
in the various environments cited before is still incomplete because, among other reasons,
characterizing variables such as the thickness of the polymer brushes on the sheared
surfaces is still difficult to accomplish [6]. These difficulties arise partly because the self –
assembly processes typically used for the modification of surfaces depend on chemical
reactions that do not necessarily yield a well – defined brush length [6]. In this regard,
computer simulations have come to play an increasingly important role, and there are now
various works investigating the role of the polymer brush length, the spacing between the
3
plates, the properties of the solvent and the influence of the electrostatic interactions, the
fluid´s shear thinning as the shear rate is increased, and the effect of changing the polymer
grafting density on the plates, to name but a few [15-24].
However, most works have been carried out for polymers in good solvent conditions , with
very few exceptions [19, 21 – 23]. It is known that some biopolymers found in an aqueous
environment are in the borderline between good solvent and theta solvent conditions [21,
25]. An important example is that of polyethylene glycol (PEG) polymer brushes on
liposomes, which are used as carriers under flow for efficient drug delivery [26, 27]. PEG is
known to have a diminishing solubility as its molecular weight increases [28], and for
molecular weights in the range of 2000 to 10000 g/mol in water it has been reported to be
slightly below its theta temperature, when placed at the human body temperature [29].
Since the mechanisms of drug delivery are of paramount importance for pharmaceutical
design, and their understanding is still far from complete, we have undertaken here the
study of the viscosity and friction of PEG brushes on surfaces that experience an external
force under theta solvent conditions, for the first time. To reach scales comparable to those
of typical non-equilibrium experiments we have carried out particle – based, mesoscopic
molecular dynamics simulations, using the method known as dissipative particle dynamics
(DPD) [30], which has been successful in predicting correctly properties of polymer
brushes both in equilibrium and non-equilibrium situations [31].
In what follows we report DPD simulations of the rheological properties of biopolymer
(PEG) brushes grafted to parallel flat plates (liposomes), including the solvent explicitly,
which is crucial to reproduce experiments measuring friction between polymer brushes [17,
18]. We want to model the behavior of tethered proteins and biopolymers in environments
4
such as vessels, where the separation between plates is essentially constant but allow for
variations in the degree of polymerization, at fixed grafting density and substrate
separation. In particular, we would like to determine the optimal conditions under which
PEG brushes on liposomes promote the flow of the latter as mechanisms for efficient drug
delivery. We obtain the viscosity and the friction coefficient as functions of the shear rate,
and degree of polymerization. To our knowledge, this is the first report of scaling behavior
for the viscosity and the friction coefficient at high shear rate of polymer brushes in theta
solvent using soft potentials. Understanding of these phenomena is useful for the
interpretation of several recent experiments in biological and other mesoscopic systems of
current academic and industrial interest.
This article is organized as follows. In the following section we introduce the interaction
model, the simulation algorithm and the details of the systems studied. In Section III the
main results are presented, accompanied by their discussion. The final section is devoted to
our major conclusions.
II. MODELS AND METHODS
We have performed DPD simulations of linear grafted polymers in the canonical ensemble
(fixed density and temperature). The DPD model is by now well known; therefore we shall
𝑪 ), that
be brief here. Three forces make up the basic DPD structure: a conservative force (𝑭⃗⃗ 𝒊𝒋
accounts for the local pressure and is proportional to an interaction constant, 𝑎𝑖𝑗 ; a
dissipative force ( 𝑭⃗⃗
𝑫 ), which represents the viscosity arising from collisions between
𝒊𝒋
particles, proportional to the relative velocity of the particles and to a constant, ; and a
5
random force (𝑭⃗⃗
𝑹 ), that models the Brownian motion of the particles, with an intensity
𝒊𝒋
given by the constant all acting between any two particles, i and j. The spatial
dependence of these forces is usually chosen as shown in equations (1-3).
𝑭⃗⃗
𝑪 = 𝑎𝑖𝑗 (1 −
𝒊𝒋
𝑟𝑖𝑗
⁄ ) 𝒆 𝒊𝒋
𝑅𝑐
𝑭⃗⃗
𝑫 = −𝛾 (1 −
𝒊𝒋
2
𝑟𝑖𝑗
⁄ )
𝑅𝑐
[𝒆 𝒊𝒋 ∙ 𝒗⃗⃗ 𝑖𝑗 ]𝒆 𝒊𝒋
𝑭⃗⃗
𝑹 = 𝜎 (1 −
𝒊𝒋
𝑟𝑖𝑗
⁄ ) 𝒆 𝒊𝒋 ξ𝑖𝑗 .
𝑅𝑐
(1)
(2)
(3)
In equations (1-3), 𝒓𝑖𝑗 = 𝒓𝑖 − 𝒓𝑗 represents the relative position vector between particles i
and j, 𝒆 𝑖𝑗 is the unit vector in the direction of 𝒓𝑖𝑗 , while 𝒗𝑖𝑗 = 𝒗𝑖 − 𝒗𝑗 is the relative
velocity between particles i and j, with 𝒓𝑖 , 𝒗𝑖 being the position and velocity of particle i,
respectively. The random variable ξij is generated between 0 and 1 with a Gaussian
distribution, zero mean and unit variance; 𝑅𝑐 is the cut off distance, beyond which all
forces are zero. The DPD beads are all of the same size, with radius 𝑅𝑐 , which is set equal
to 1. The constants in the dissipation and random forces are not independent, and satisfy the
relation [32]:
𝜎 2
2𝛾
= 𝑘𝐵 𝑇
(4)
which, in effect, fixes the temperature; here, kB is Boltzmann’s constant. This natural
thermostat that arises from the balance between the dissipative and random forces is a
defining feature of the DPD model [31]. Also, the short-range nature of the DPD forces,
and their linearly decaying spatial dependence, allow the use of relatively large time steps
6
when integrating the equation of motion. The DPD particles are representations of sections
of fluid rather than physical particles, and can group several atoms or molecules, making
the DPD method an attractive alternative to study systems at the mesoscopic level. Note
also that the forces in equations (1)-(3) obey Newton’s third law, which means momentum
is conserved locally, and globally, which in turn preserves any hydrodynamic modes
present in the fluid. This is a feature of fundamental importance when studying non
equilibrium properties of fluids, since loss of information about these modes can lead to a
different phase from that obtained when they are fully accounted for [33]. The DPD
interaction model has been used successfully to predict equilibrium properties of polymer
melts [34], surfactants in solution [35], and colloidal stability [36], among others. For
further reading, see reference [31].
All our simulations are performed in reduced units (marked with asterisks); distances are
reduced with the cutoff radius, Rc, which for a coarse – graining degree equal to 3 water
molecules per DPD particle is equal to Rc,=6.46 Å [31], hence r=r*Rc. The time step t is
2 𝑘𝐵𝑇⁄
reduced with 𝛿𝑡 = (𝑚𝑅𝐶
)1 2⁄ 𝛿𝑡 ∗ , where m is the mass of a DPD particle, while
energy is reduced with kBT. All other quantities can be reduced through combinations of the
3⁄
these relations, for example, the viscosity: 𝜂 = 𝜂 ∗(𝑘𝐵𝑇𝛿𝑡 𝑅𝐶
). Using the mass of 3 water
molecules per DPD particle, at room temperature, one obtains 𝛿𝑡 ≈ (6.3 × 10−12s)𝛿𝑡 ∗ and
∗ = 1, t*=0.03. In
𝜂 ≈ (9.64 cP)𝜂∗ . All our simulations are performed for 𝑘𝐵𝑇 ∗ = 𝑚∗ = 𝑅𝐶
addition to the fluid monomeric particles, our system contains soft parallel surfaces as
models for biological membranes, and polymer chains attached to them, forming brushes.
The polymers are built as linear chains of DPD beads joined by freely rotating harmonic
2⁄
springs [37]. The spring constant is chosen as 𝜅 = 𝜅 ∗(𝑘𝐵𝑇 𝑅𝐶
), where 𝜅 ∗ =100 and the
7
∗ 𝑅𝐶 , with 𝑟𝑒𝑞
∗ = 0.7 in all cases [38]. In our
spring’s equilibrium position as 𝑟𝑒𝑞 = 𝑟𝑒𝑞
simulations we have fixed the distance between the plates, which are placed perpendicular
to the z-direction, at a distance of D*=7. Periodic boundary conditions are applied on the xy
– plane but not in the z – direction, to reinforce the confinement. The equation of motion is
solved using the velocity Verlet algorithm, adapted to DPD [39]. The parameters of the
dissipative and random forces intensities are, respectively, = 4.5 and = 3.0, so that kBT*
= 1 (see equation (4)). The value of the conservative force intensity (see equation (1)) was
set to aij = 78.0 for all cases, namely for the interaction between particles of the same type
(solvent – solvent, monomer – monomer) and for particles of different type (solvent –
monomer). The choice aij = 78.0 is obtained when one uses a coarse graining degree that
groups 3 water molecules in a single DPD bead [35]. Since all particle – particle
interactions are equal, the polymers are in a theta – solvent, as is the case for PEG –
covered liposomes in the human circulatory system [29]. It has been shown that this model
for PEG on colloidal surfaces can successfully predict brush scaling laws [27], as well as
adsorption isotherms on metallic oxides and disjoining pressure profiles [36]. The soft
membranes on which these biopolymers are tethered are modeled as linearly decaying
forces that act on the ends of the simulation box (in the z – direction), given by the
following expression:
𝐹𝑤 (𝑧𝑖 ) = 𝑎𝑤 [1 −
𝑧𝑖
𝑧𝐶
].
(5)
In equation (5), 𝐹𝑤 (𝑧𝑖) symbolizes the force exerted by the effective wall on particle i,
whose position component in the z – direction is zi. The intensity of the force is given by
the constant 𝑎𝑤 , while zC is the cutoff distance set also equal to 1, which defines the reach
8
of this force, i.e., 𝐹𝑤 (𝑧𝑖) = 0 for zi > zC. Equation (5) represents a soft membrane because
the maximum repulsion between the beads and the surface is 𝑎𝑤 when zi =0; harder walls
can be used for other DPD applications [36]. It is known that liposomes can be up to about
3 orders of magnitude larger than their brush thickness, and the solvent [2, 40], therefore
we consider planar walls only. To fix one end of the polymers on these substrates we chose
their interaction equal to aw=70.0 (reduced DPD units), while the rest of the polymer beads
and the solvent interact with the walls with an intensity given by aw=140.0. With this
choice of parameters, the polymer end experiences a less repulsive force toward the walls
than the rest of the polymer (and solvent) beads, and is therefore adsorbed on them. The
tethered ends of the polymers remain free to move on the xy – plane, of course, as occurs
also for biopolymers interacting with membranes. Although this model for membrane is
soft, it remains impenetrable to polymer and solvent molecules, and we assume moreover
that the drug molecules have already been incorporated into the structure.
To carry out non equilibrium simulations of these systems, we apply a constant shear
velocity to the polymer ends adsorbed on the walls, of equal magnitude but opposite sign
for beads on different plates. This is equivalent to moving the plates in opposite direction
under the influence of a fixed external force that keeps the plates moving with constant
speed, known as Couette flow [19 – 23], see Figure 1. Some of the solvent monomers are
seen to penetrate the PEG brushes and carry them along in their flow to produce
lubrication, while there is no interpenetration of the brushes. As pointed out before, our
simulations are performed in the canonical ensemble, where particle density and
temperature are kept constant; previous works have shown that completely equivalent
results are obtained if one performs the simulations in the grand canonical ensemble, at
9
fixed chemical potential, volume and temperature [17]. To keep the simulations results
invariant to the particular choice of aij interaction parameters, we fix the global
dimensionless density to 3 [35] in all the simulations reported here; this means that when
the degree of polymerization of the PEG brushes (N) is increased, the number of solvent
molecules is reduced.
Figure 1. Snapshot of two linear PEG brushes made up of N=7 beads. The lateral dimensions of the
∗ =35, and the spacing between the plates is 𝐿𝑧
∗ =35, 𝐿𝑦
∗ = 𝐷 ∗=7. The grafting
simulation box are 𝐿𝑥
density is equal to =0.3. A constant shear velocity of magnitude v0
*=1.0 is applied to the tethered
beads of each polymer (in dark blue); the rest of the polymer beads are shown in yellow. The
solvent monomers are shown in red. Notice how the solvent penetrates the polymer brushes all the
way up to the surfaces for this grafting density. See text for details.
The grafting density is defined as =Np/A*, where Np is the number of polymer chains
∗ 𝐿𝑦
∗ is its reduced area. Our simulations are performed at
tethered on the surface and A*=𝐿𝑥
grafting densities that lie within the brush regime (, where N is the degree of
10
polymerization, see [27]), where the “stealth” properties of the PEG brushes work best [41].
We obtain the viscosity (), and the friction coefficient () of the fluid, through the
relations [19]
𝜂 =
〈𝐹𝑥 (𝛾 )〉 𝐴⁄
𝛾
,
𝜇 =
〈𝐹𝑥 (𝛾 )〉
〈𝐹𝑧 (𝛾 )〉
.
(6)
(7)
In equations (6) and (7) above, 〈𝐹𝑥 (𝛾 )〉 and 〈𝐹𝑧(𝛾 )〉 are the mean forces that the particles on
the surfaces experience along the flow direction, and perpendicularly to it, respectively; the
brackets indicate an ensemble average. Equation (6) can be understood as the local
definition of viscosity, applied to the entire sample. Considering a liquid confined between
two planar walls, a linear flow can be generated by moving for example the top wall at
constant velocity. This motion requires that a steady force be applied on the top wall, and
an equal in magnitude but opposite in direction force, on the bottom wall, see Figure 1. In
linear flow, the velocity gradient (shear rate, 𝛾 ) is constant throughout the liquid. The shear
∗ 𝐷∗⁄
rate 𝛾 is defined as 2𝑣0
, where v0
* is the flow velocity exerted on the grafted
monomers, and D* is the surface separation (see Figure 1). To obtain first a local definition
of viscosity, a small cubic volume in the fluid (fluid element) can be considered. This fluid
element undergoes strain, being deformed from a cubic to a parallelepiped shape, with
increasing deformation in time. The amount of strain is the added lateral displacement of
top and bottom planes of the fluid volume divided by its height. This strain increases at
constant rate in time for linear flow, such that the strain rate i.e. the strain per unit time, is
the velocity gradient in the limit of an infinitesimal fluid element. The strain rate of
deformation is therefore the shear rate, 𝛾 . On the other hand, a force acts on the top and
11
bottom planes of the fluid element, with identical magnitude and opposite direction. The
center of mass of the fluid element, suffers no acceleration and therefore the total external
force on it must vanish. These forces must be proportional to the area of the top and bottom
planes of the fluid element. Also, the force-per-unit-area or stress, , must be uniform from
top to bottom in the fluid element since the fluid is not accelerating. For linear flow, this is
true not only for a fluid element but for the whole fluid: every part of the fluid is under the
same stress, which produces everywhere the same velocity gradient (𝛾 ). The definition of
viscosity assumes that the shear rate (𝛾 ) produced in a fluid element is proportional to the
shear stress () exerted on it. The viscosity is defined as the constant of proportionality
between them: = 𝛾 , or = 𝛾 . The total shear stress on the sample () can be
determined from the average total force on the brush heads divided by the wall area (Fx/A)
while the shear rate (𝛾 ) is extracted from the slope of the linear fit of the average velocity
profile. The viscosity is then expressed by equation (6). As for the friction coefficient,
shown in equation (7), its calculation follows directly from the analogy with the friction
coefficient between solid surfaces, namely Fx=Fz, where Fz is the force acting
perpendicularly to the surface, and Fx is the force acting parallel to it, which is responsible
for the shear stress ( = Fx/A) [24]. Our results were obtained from averages of simulations
of up to 4×103 blocks, with each block composed of 2×104 time steps, using the first 2×105
time steps for equilibration and the rest for the production phase; when properly
dimensionalized this represents a time observation window of 0.12 ms. A typical density
profile of the solvent and polymer monomers is shown in Figure 2, where it is clear that the
solvent penetrates the PEG brushes and reaches the surfaces, at that given grafting density.
12
Figure 2. Density profiles of the PEG brushes (dashed line, in blue) with polymerization degree
N=7, and the solvent (full line, in black). The polymer monomers order near the walls, which leads
to their structuring, represented here by peaks near the walls, while there is a relatively free (bulk
like) fluid made up of solvent molecules at the center of the simulation box. For the case shown in
∗ =𝐿𝑦
∗ =7, D*=7, v0
this figure, 𝐿𝑥
*=1.0 and =1.0. Note there is very little interpenetration between
the brushes. The brush structuring is rather strong because the chains are relatively short.
The profiles of the opposing polymer brushes show very little interpenetration and an
almost bulk like concentration of solvent monomers at the center of the simulation box (see
Figure 2), which reduces the viscosity of the system. The structuring of the PEG brush
monomers, represented by the maxima in Figure 2, coincides with that of the solvent
monomers, indicating that even for a relatively large grafting density as that shown in the
figure, the solvent is able to reach the membrane. The maxima of both profiles (the
solvent’s and the brushes’) occur at the same positions because the interactions between
13
0123456702468 (z)z solvent brushthem are the same (recall aij=aii=78.0 for all i and j), as it is befitting for theta – solvent
conditions. The strong layering is expected when the chains are relatively short, as in
Figure 2, because in such case monomers are more easily arranged that when long chains
form the brushes [16-20]. In Figure 3 we present the velocity profile for PEG brushes of
polymerization degree N=14 and grafting density =0.30 under a shear velocity equal to
v0
*=0.1. Clearly, there appears a linear gradient in the center of the channel formed between
∗ 𝐷 ∗⁄ ,
liposomes, from which the shear rate 𝛾 can be obtained, through the relation 𝛾 = 2𝑣0
as pointed out before. The inset in Figure 3 shows the temperature profile of the complex
fluid in the z-direction, which indicates the brushes are at the temperature fixed by the
thermostat (kBT*=1), although at large shear rates one expects the brush to “heat up”
somewhat [19], as a consequence of the increased dissipation rate.
14
Figure 3. Velocity profile at the center of the channel between liposomes for PEG brushes with
N=14 and =0.30, using v0
*=0.1. The box size is Lx
*=Ly
*=7, D*=7. The inset shows the temperature
profile (Tx
*) in the z-direction. All quantities are reported in reduced units.
Most simulations were performed for a brush grafting density *=0.30, except where
indicated otherwise, because for this value the average distance between grafted heads on
the surface is smaller than their radius of gyration in theta solvent, for all values of N we
studied [27]. By doing so one makes sure that the grafted polymer chains are in the brush,
rather than the mushroom regime [27], which is important for the situation we are
interested in modeling. The polymerization degree was varied in the range N=1 up to N=25,
so that scaling laws could be extracted from the data. Finally, the shear rate was chosen to
vary from 𝛾 = 0 to 𝛾 = 0.30, except where indicated otherwise, so that we could compare
our results with those available in the literature [16-20]. Molecular dynamics simulations
like ours solve the equation of motion essentially exactly [33], although the use of soft
forces in DPD (see, for example, equation (1)) might limit its applicability to situations
where atomistic detail is not important, as is the case for the nanotribology studies that are
the focus of this work.
III. RESULTS AND DISCUSSION
One would like to simulate systems with a large number of particles so that realistic
situations can be reproduced, while simultaneously keeping such number manageable to
obtain results in a timely fashion. Therefore, we have carried out simulations to quantify the
extent of finite size effects in the viscosity of the polymer brushes, if any. Figure 4 shows
the mean viscosity of the PEG brushes and the solvent, calculated with equation (6), as a
15
function of the lateral size of the membranes (Lx
*=Ly
*), at fixed separation between them
(D*=7), grafting density (=0.30) and shear rate (𝛾 =0.28). The results in Figure 4 cover a
range that goes from 103 particles up to 105 particles while the relative change in the
viscosity amounts to less than 1 %, showing that finite size effects are not important in the
calculation of the viscosity, which allows us to make correct predictions using relatively
small systems.
Figure 4. Finite size effects in the viscosity, . The symbol Lx
* represents the size of the simulation
box in the x – direction, which is equal to that in the y – direction, Ly
*. In all cases, Lz
*=D*=7, the
shear rate is 𝛾 = 0.28, and the grafting density is =0.30. All quantities are reported in reduced
units. Lines are only guides for the eye.
16
010203040506070801.4801.4851.4901.4951.5001.5051.5101.5151.520 Lx* Similar results have been found in equilibrium DPD simulations [42] which proved that
finite size effects in the interfacial tension between two model fluids were as small as those
found in the viscosity and shown in Figure 4, when the appropriate ensemble was used.
This behavior is to be attributed to the soft, short – range repulsive nature of the DPD
forces (see equations (1)-(3)), although these effects may become non negligible if a long –
range interaction is included, such as the electrostatic one. In the remainder of this work we
present results obtained using only the smallest simulation box in Figure 4.
Let us now proceed to discuss the results obtained for the mean viscosity of a complex fluid
made up of PEG brushes and an aqueous solvent, as a function of the shear rate for
different degrees of polymerization (N), as seen in Figure 5. The simulations were
performed with a cubic box of lateral size L*=7 and for =0.30. For very small values of 𝛾
(less than 10-4) the calculation of requires of very long simulations due to poor signal –
to – noise ratios [19], which makes it very difficult to obtain accurately the zero – shear
viscosity, 0, from simulations. However, a good estimate of it can be obtained from the
extrapolation of the rheology profiles in Figure 5 as 𝛾 approaches zero. For the shorter PEG
brushes (N=1 and N=3) the fluid displays essentially Newtonian behavior, whereas for the
case with N=7 the fluid shows shear – thinning. One observes also that as N is increased so
is the extrapolated 0 value, which is a consequence of the combined effects of an
increasing brush thickness and a reduced number of solvent particles (to keep the global
density =3). Increasing N is equivalent to reducing the separation between the
membranes, D*, when the grafting density is kept constant, and yields increasing values of
0 [18]. Hence, we have opted to vary N rather than D* in this work. The simulations shown
in Figure 5 were carried out for relatively short polymers with the purpose of establishing
17
the minimum polymerization degree at which the fluid started to show shear thinning,
which, from Figure 5, is N=7. For polymerization degree smaller than 7, it is difficult for
the polymer chains to interact with the solvent monomers enough to reduce the viscosity as
the shear rate is increased, hence the fluid remains approximately Newtonian (N=1 and N=3
in Figure 5).
Figure 5. Viscosity for a fluid made of polymer brushes of various degrees of polymerization (N)
and monomeric solvent particles. Notice how the shorter brushes show virtually no shear – thinning,
and behave almost like Newtonian fluids. In all cases, Lx
*=Ly
*=D*=7, and the grafting density is
=0.30. All quantities are reported in reduced units. Lines are only guides for the eye.
It has been reported that increasing the polymerization degree of PEG brushes on liposomes
increases their circulation longevity [28], which would be a desirable aspect to improve
18
drug delivery mechanisms. However, as its degree of polymerization grows, PEG’s
solubility in aqueous environment is reduced [29], hence there would be a competition
between these trends that can be investigated with DPD simulations, varying N. Also, it is
important to find out if the viscosity behaves qualitatively the same as a function of shear
rate, regardless of the value of N, so that general conclusions are obtained that can be used
to interpret a variety of experiments [24]. Before presenting our results for the effective
mean viscosity of systems where PEG brushes have larger polymerization degrees we
comment on the behavior of the mean shear stress on the membranes along the direction of
the flow as a function of the shear, 〈𝐹𝑥 (𝛾 )〉. Following Galuschko and coworkers [43], we
find a critical shear rate for our systems, 𝛾 ∗ , as the value of the shear rate when the behavior
of 〈𝐹𝑥 (𝛾 )〉 changes from linear to sub linear, which depends on N, being 𝛾 ∗~8.6 × 10−3 for
N=14, and 𝛾 ∗~2.9 × 10−3 for N=25. Finding 𝛾 ∗ is tantamount to finding the Weissenberg
number, We, since 𝑊𝑒 = 𝛾 𝛾 ∗⁄ , and it signal when shear – thinning behavior starts to set in
(We ≥ 1). Figure 6 shows the dependence of the normalized shear stress, 𝑢 ≡
〈𝐹𝑥 (𝛾 )〉 〈𝐹𝑥 (𝛾 ∗)〉
⁄
, on We for fluids with brushes with polymerization degrees equal to
N=14, 20 and 25. The data for the three brushes are seen to collapse reasonably well,
especially at low We; this is the so called linear regime, 𝑢~𝑊𝑒, (dashed blue line in Figure
6). For larger values of the Weissenberg number, we find that the power law 𝑢~𝑊𝑒 𝜅 is
obeyed with 𝜅 = 0.69 (solid black line in Figure 6) for our systems under theta – solvent
⁄
conditions. Galuschko et al. [43] predict 𝜅 = 9 13 ≈ 0.69
using scaling arguments for
what they call “dry” polymer brushes, which are brushes where the hydrodynamic and
excluded volume interactions are screened out , for fairly concentrated solutions and
polymer melts, where the corresponding Flory exponent is =0.5 [44]. Our simulations
19
predict a value in excellent agreement with scaling arguments [43] because under theta –
solvent conditions and at the concentrations we modeled, the brushes are almost
indistinguishable from the solvent, and act as a melt.
Figure 6. Normalized mean shear stress (u) on the surfaces in the direction of the flow as a function
of the Weissenberg number, We, for PEG brushes of three different polymerization degrees, N. In
all cases, Lx
*=Ly
*=D*=7, and the grafting density is =0.30. The dashed blue line represents the
linear regime, 𝑢~𝑊𝑒 (We < 1), and the solid black line the power law 𝑢~𝑊𝑒 𝜅 , with 𝜅 = 0.69. See
text for details.
Recent molecular dynamics simulations [45] of strongly compressed polyelectrolyte
brushes under shear at melt concentrations find that the shear force 〈𝐹𝑥 (𝛾 )〉 that appears in
equations (6) and (7) behaves at high shear rate (𝛾 ) as 〈𝐹𝑥 (𝛾 )〉~𝛾 0.69 , in agreement with our
predictions in Figure 6, and with the scaling arguments of Galuschko et al. [43]. Although
20
10-210-110010110210-210-1100101 N=14 N=20 N=25uWethere are explicit electrostatic interactions in the work of Spirin and Kreer [45], they argue
that their brush and solvent system behaves as a melt of neutral polymers at high We
because of the immobilization of the counterions. Therefore, the scaling exponent =0.69
we obtained in Figure 6 is very robust.
The effective mean viscosity data of these systems can also be collapsed as a function of
We if one defines 𝑠 = 𝜂 𝜂0⁄ [43], for We > 1, as seen in Figure 7(a). The value of the zero
shear viscosity 𝜂0 was obtained from extrapolation, using the same procedure as the one
described in the discussion of Figure 5. All three cases experience considerable shear
thinning as the Weissenberg number is increased, and in this regime the data follow the
power law 𝑠~𝑊𝑒 𝜁 , with 𝜁 = −0.31, see the black line in Figure 7(a). To our knowledge,
this is the first time such scaling law has been obtained in theta solvent.
In reference [43] the value 𝜁 = −0.43 was obtained for linear Lennard – Jones brushes in
good solvent. A similar value ( 𝜁 = −0.42 ) was found in simulations of bottle – brush
polyelectrolytes in good solvent [18]. Hence, polymer architecture appears to be of
secondary importance, compared with the quality of the solvent or the polymer
concentration. Galuschko and collaborators [43] proposed also the following relation
between 𝜅 and 𝜁 :
𝜅 − 𝜁 = 1
(8)
which is clearly fulfilled in our theta solvent simulations. Therefore the relation between
the scaling exponents is the same regardless of solvent quality, and is more fundamental
than the architecture of the polymer chains or the particular values of the exponents
themselves, separately, as is known to be the case for equilibrium scaling exponents [46].
21
22
10-210-110010110210-1100 sWe N=14 N=20 N=25s~We -0.31(a)101102103101102. RBC(mPa s)(s-1).(b)Figure 7. (a) Normalized effective mean viscosity 𝑠 = 𝜂 𝜂0⁄ as a function of the Weissenberg
number, We, for fluids with PEG brushes of three different polymerization degrees, N. In all cases,
Lx
*=Ly
*=D*=7, and the grafting density is = 0.30. The line represents the power law, 𝑠~𝑊𝑒 𝜁 ,
with 𝜁 = −0.31. (b) Viscosity of a fluid made up of red blood cells (RBC) dispersed in a 40 kDa
liposome – dextran medium, as a function of the shear rate [47]. The blue line represents the power
law 𝜂~𝛾 −0.31. See text for details.
In Figure 7(b) we have included the high shear rate values of the viscosity of human red
blood cells (RBC) in a liposomal suspension [47], which shows evidently shear – thinning.
This non – Newtonian behavior is interpreted as being the result of the dissociation of
flocculated liposomes at the higher values of the shear rate, an interpretation that is
consistent with our predictions in Figure 7(a). The blue line in figure 7(b) represents the
power law we have obtained for PEG – grafted surfaces in a theta solvent, which models
accurately the results for RBC at high values of the shear rate, providing confirmat ion for
our predictions. Experiments carried out for gels used as vehicles for drug delivering
liposomes [48], such as hydroxyethylcellulose (HEC) and a mixture of HEC and an acrylic
acid – based polymer (carbopol 974) yield values of the viscosity at high shear rates very
close to that found in Figure 7(a), namely 𝜁 = 0.30 ± 0.01 for HEC and 𝜁 = 0.310 ± 0.009
for the mixture. The fact that the exponent in these experiments is found to be very close to
our prediction suggests that it is of paramount importance to take into account the
properties of the solvent (theta conditions), as well as those of the brush (PEG) , and lends
additional support to these calculations.
Adding polymer brushes to sheared surfaces not only modifies the viscosity of the system
but it has been shown to reduce its friction coefficient [10] also. Since this is a parameter
23
that can be directly compared with experiments, we have obtained the friction coefficient of
PEG brushes in an explicitly included, theta solvent. Using equation (7) one calculates the
mean friction coefficient between the two sliding surfaces. The data in Figure 8(a) show
that the friction coefficient at constant PEG grafting density, increases with the shear rate,
in agreement with trends found by others [18, 43, 49]; it increases also with N because the
overlap between the brushes also grows. Moreover, the values for PEG brushes in theta
solvent are somewhat larger than those obtained in simulations of linear brushes in good
solvent [17]; this is of course expected because the brush thickness is reduced when it is
immersed in a theta solvent, and the contacts between the polymers grow when going from
good to theta solvent [2]. Experiments by Klein and collaborators [50] on the friction of
polymer brushes in theta solvent compared to good solvent show that the friction
coefficient can be larger in the former by up to three orders of magnitude than in the latter.
The reason relies on the fact that in theta solvent conditions the brushes can interdigitate
and the solvent can penetrate the brushes (see Figure 2), reducing the freely flowing solvent
in the center of the channel, thereby increasing the shear force and the friction. Klein and
coworkers [50] obtain values for which are close to our predictions, shown in Figure 8(a).
When electrostatic interactions are included the values for obtained from computer
simulations turn out to be slightly larger and closer to 1 for good solvent conditions at high
values of the shear rate [18, 49], due to the added osmotic pressure of the ions. However, in
biological systems like those modeled here, the quality of the solvent plays a key role in
providing a mechanism that promotes a lower viscosity environment.
24
25
0.000.050.100.150.200.250.300.000.050.100.150.200.250.300.35 N=25 N=20 N=14 N=7(a).10-210-110010110210-210-110010110-410-310-210-110010-410-310-210-1 N=25 N=20 N=14*We(b). N=7Figure 8. (a) Friction coefficient for polymer brushes of various degrees of polymerization (N),
obtained using equation (7). In all cases, Lx
*=Ly
*=D*=7, and the grafting density is = 0.30. Lines
are only guides for the eye. (b) The friction coefficient data shown in (a), normalized by its value at
𝛾 ∗ for the largest 3 brushes, *, as a function of the Weissenberg number, 𝑊𝑒 = 𝛾 𝛾 ∗⁄ . The black
line represents the fit to the power law 𝜇 𝜇∗⁄ ~𝑊𝑒 𝜅 , with =0.69. The inset shows the friction
coefficient for the polymer with the smallest degree of polymerization (N=7, pink rhombi), with the
blue line representing a linear fit, 𝜇~𝛾 .
In Figure 8(b) we show the same data as in Figure 8(a), on logarithmic scales to emphasize
power law behavior. For the shortest brush (see inset in Figure 8(b)), made up of chains
with polymerization degree equal to N=7, there is a linear relation between the friction
coefficient and the shear rate, 𝜇~𝛾 , for the entire range of shear rate values used in the
simulations (blue line). To see sublinear behavior here one would have to impose larger
shear. For the other polymerization degrees shown (N=14 to 25) the main panel in Figure
8(b) shows that the data collapse rather well on a single curve when the friction coefficient
() is normalized by its value (*) at the start of the shear – thinning regime (𝛾 ∗ ). The black
line is the fit to the power law 𝜇 𝜇∗⁄ ~𝑊𝑒 𝑘 , with =0.69. This is the same exponent as the
one obtained from the normalized shear stress data, in Figure 6, indicating that the force
〈𝐹𝑧(𝛾 )〉 (see equation (7)) is essentially independent of shear once the brushes have reached
a certain length. At the same time, the fact that the exponent is =0.69 at 𝑊𝑒 > 1 in all
three cases (N=14, 20 and 25) shows that 〈𝐹𝑥 (𝛾 )〉 (see equation (7)) is responsible for the
behavior of the friction coefficient in the high shear rate regime. The same behavior is
found for Lennard – Jones brushes [43], namely that the friction coefficient dependence on
26
We scales with same exponent as that of the shear stress, although in such case =0.57
because that system is under good – solvent conditions.
IV. CONCLUSIONS
We have shown that complex fluids with polymer brushes under theta solvent conditions
display rheological characteristics that differ in detail from their good solvent counterparts,
but that obey nevertheless the same general scaling properties, in particular for the shear
stress and the viscosity. The exponent for the shear stress as a function of We obtained from
our simulations, =0.69, is in excellent agreement with the value predicted using scaling
arguments [43] for dense polymer brushes where excluded volume interactions are
screened out, as in our model. The viscosity we obtained scales with the shear rate with an
exponent equal to 𝜁 = −0.31 , which reproduces remarkably well measurements of the
viscosity of red blood cells dispersed in a liposome carrying aqueous fluid at high shear rate
[47], as well as other experiments [48]. The friction coefficient data as a function of shear
rate for different polymerization degrees of the chains making up the brushes were found to
collapse on a universal curve whose behavior at We > 1 follows the same power law as the
shear force, i.e., 𝜇~𝑊𝑒 𝜅 , with =0.69, in agreement with trends found in simulations under
good – solvent conditions [43], albeit with a value of particular to those conditions. Ours
are the first simulations, to the best of our knowledge, of the scaling of viscosity and the
friction coefficient for systems under theta – solvent conditions.
It is argued that our simulations are useful for understanding the behavior of biopolymer
brushes coating drug – carrying liposomes in an aqueous environment that acts as a theta
27
solvent, under shear. Increasing the thickness of the brush through the degree of
polymerization or the shear velocity on the surfaces is shown to raise the friction
coefficient, which is a deleterious effect for the transport properties of these liposomes.
However, when liposomes are covered by PEG bushes in a theta solvent at high shear rates,
their flowing characteristics make them optimal carriers for drug delivery, as the polymer
brushes imprint them with efficient injectable characteristics (low viscosity at high shear
rate) while at the same time providing them with thermodynamically stable (“stealth”)
mechanisms. These simulations have the additional advantage of including hydrodynamic
interactions, as well as the solvent explicitly, and being mesoscopic they reproduce length
and time scales comparable with those of environments of biological interest. Therefore, we
believe our work should be useful in the improved design of drug – carriers, the rheological
characterization of sheared brushes, and in the establishing of general scaling laws for non
– equilibrium polymer brushes.
V. ACKNOWLEDGEMENTS
The authors would like to thank ABACUS, CONACyT grant EDOMEX-2011-C01-
165873, for funding. AGG thanks M. A. Balderas Altamirano, C. Carmín, E. de la Cruz, J.
P. López Neria and E. Pérez for instructive discussions.
VI. REFERENCES
28
[1] D. H. Napper, Polymeric Stabilization of Colloidal Dispersions, Academic Press,
London, 1983.
[2] J. N. Israelachvili, Intermolecular and Surfaces Forces, 3rd. Ed., Academic Press, New
York, 2011.
[3] A. Mollaei, B. Maini, J. Can. Petroleum Tech. 49, 65 (2010).
[4] M. A. Cohen Stuart, W. T. S. Huck, J. Genzer, M. Müller, C. Ober, M. Stamm, G. B.
Sukhorukov, I. Szleifer, V. V. Tsukruk, M. Urban, F. Winnik, S. Zauscher, I. Luzinov, S.
Minko, Nature Materials, 9, 101 (2010).
[5] S. Iyer, R. M. Gaikwad, V. Subba – Rao, C. D. Woodworth, I. Sokolov, Nature
Nanotech. 4, 389 (2009).
[6] R. Advincula, W. J. Brittain, K. C. Caster, J. Rühe (eds.), Polymer Brushes, Wiley-
VCH, 2004.
[7] P. G. de Gennes, Macromolecules 13, 1069 (1980).
[8] S. Alexander, J. Phys. (Paris) 38, 983 (1977).
[9] S. T. Milner, T. A. Witten, M. E. Cates, Europhys. Lett. 5, 413 (1988).
[10] J. Klein, E. Kumacheva, D. Mahalu, D. Perahia, L. J. Fetters, Nature 370, 634 (1994).
[11] L. Han, D. Dean, P. Mao, C. Ortiz, A. J. Grodzinsky, Biophys. J. 92, 1384 (2007).
[12] J. Klein, Science 323, 47 (2009).
[13] S. Weinbaum, J. M. Tarbell, E. R. Damiano, Annu. Rev. Biomed. Engng. 9, 121
(2007).
[14] A. Samal, Y. Sultana, M. Aqil, Curr. Drug Delivery 4, 297 (2007).
[15] P. Y. Lai, C. Y. Lai, Phys. Rev. E 54, 6958 (1996).
[16] T. Kreer, M. H. Müser, K. Binder, J. Klein, Langmuir 17, 7804 (2001).
[17] F. Goujon, P. Malfreyt, D. J. Tildesley, Mol. Phys. 103, 2675 (2005).
29
[18] J. M. Carrillo, W. M. Brown, A. V. Dobrynin, Macromolecules 45, 8880 (2012).
[19] C. Pastorino, K. Binder, T. Kreer, M. Müller, J. Chem. Phys. 124, 064902 (2006).
[20] F. Goujon, P. Malfreyt, D. J. Tildesley, Soft Matter 6, 3472 (2010).
[21] M. Deng, X. Li, H. Liang, B. Caswell, G. E. Karniadakis, J. Fluid Mech. 711, 192
(2012).
[22] D. Irfachsyad, D. J. Tildesley, P. Malfreyt, Phys. Chem. Chem. Phys. 4, 3008 (2002).
[23] C. Pastorino, K. Binder, M. Müller, Macromolecules 42, 401 (2009).
[24] K. Binder, A. Milchev, J. Pol. Sci. Part B: Pol. Phys. 50, 1515 (2012).
[25] F. Alarcón, E. Pérez, A. Gama Goicochea, Eur. Biophys. J. 42, 661 (2013).
[26] D. D. Lasic, F. Martin (eds.), Stealth Liposomes, CRC Press, Florida, 1995.
[27] P. L. Hansen, J. A. Cohen, R. Podgornik, V. A. Parsegian, Biophys. J. 84, 350 (2003).
[28] C. Allen, N. Dos Santos, R. Gallagher, G. N. C. Chiu, Y. Shu, W. M. Li, S. A.
Johnstone, A. S. Janoff, L. D. Mayer, M. S. Webb, and M. B. Bally, Bioscience Reports 22,
2, (2002).
[29] C. Ö. Dinç, G. Kibarer, A. Güner, J. Appl. Polym. Sci. 117, 1100 (2010).
[30] P. J. Hoogerbrugge and J. M. V. A. Koelman, Europhys. Lett. 19, 155 (1992).
[31] T. Murtola, A. Bunker, I. Vattulainen, M. Deserno, M. Karttunen, Phys. Chem. Chem.
Phys. 11, 1869 (2009).
[32] P. Español and P. Warren, Europhys. Lett. 30, 191 (1995).
[33] D. Frenkel and B. Smit, Understanding Molecular Simulation, Academic, New York,
2002.
[34] R. D. Groot, J. Chem. Phys. 118, 11265 (2003).
[35] R.D. Groot and P.B. Warren, J. Chem. Phys. 107, 4423 (1997).
[36] A. Gama Goicochea, Langmuir 23, 11656 (2007).
30
[37] M. Murat and G. S. Grest, Phys. Rev. Lett. 63, 1074 (1989).
[38] A. Gama Goicochea, M. Romero-Bastida, R. López-Rendón, Mol. Phys. 105, 2375
(2007).
[39] I. Vattulainen, M. Karttunen, G. Besold, J. M. Polson, J. Chem. Phys. 116, 3967
(2002).
[40] D. D. Verma, S. Verma, G. Blume, A. Fahr, Intl. J. Pharm. 258, 141 (2003).
[41] D. Needham, K. Hristova, T. J. McIntosh, M. Dewhirst, N. Wu, D. D. Lasic, J.
Liposome Res. 2, 411 (1992).
[42] M. E. Velázquez, A. Gama Goicochea, M. González-Melchor, M. Neria, J. Alejandre,
J. Chem. Phys. 124, 084104 (2006).
[43] A. Galuschko, L. Spirin, T. Kreer, A. Johner, C. Pastorino, J. Wittmer, J. Baschnagel,
Langmuir 26, 6418 (2010).
[44] P. G. de Gennes, Scaling Concepts in Polymer Physics, Cornell University Press, New
York, 1979.
[45] L. Spirin and T. Kreer, ACS Macro Lett. 2, 63 (2013).
[46] N. Goldenfeld, Lectures on Phase Transitions and the Renormalization Group,
Addison Wesley Publishing Company, 1992.
[47] H. Sakai, A. Sato, N. Okuda, S. Takeoka, N. Maeda, E. Tsuchida, Am. J. Physiol.
Heart Circ. Physiol. 297, 583 (2009).
31
[48] S. Mourtas, S. Fotopoulou, S. Duraj, V. Sfika, C. Tsakiroglou, S. G. Antimisiaris, J.
Coll. Interface Science 317, 611 (2008).
[49] J. M. Carrillo, D. Russano, A. Dobrynin, Langmuir 27, 14599 (2011).
[50] J. Klein, E. Kumacheva, D. Perahia, D. Mahalu, S. Warburg, Faraday Discuss. 98,
173 (1994).
32
|
1612.03146 | 1 | 1612 | 2016-12-07T23:52:17 | Scattering and Absorption Control in Biocompatible Fibers towards Equalized Photobiomodulation | [
"physics.bio-ph",
"physics.optics"
] | Transparent tissue scaffolds enable illumination of growing tissue to accelerate cell proliferation and improve other cell functions through photobiomodulation. The biphasic dose response of cells exposed to photobiomodulating light dictates that the illumination to be evenly distributed across the scaffold such that the cells are neither under nor over exposed to light. However, equalized illumination has not been sufficiently addressed. Here we analyze and experimentally demonstrate spatially equalizing illumination by three methods, namely, engineered surface scattering, reflection by a gold mirror, and traveling-waves in a ring mesh. Our results show that nearly equalized illumination is achievable by controlling the light scattering-to-loss ratio. This demonstration opens opportunities for dose-optimized photobiomodulation in tissue regeneration. | physics.bio-ph | physics | Scattering and Absorption Control in Biocompatible
Fibers towards Equalized Photobiomodulation
J. George,1 H. Haghshenas,1 D. D'Hemecourt,1 W. Zhu2, L. Zhang2, V. Sorger1,*
1Department of Electrical and Computer Engineering, The George Washington University, Washington, D.C. 20052, USA
2Department of Mechanical and Aerospace Engineering, The George Washington University, Washington, D.C. 20052,
USA
*Corresponding author: [email protected]
Abstract:
Transparent tissue scaffolds enable illumination of growing tissue to accelerate cell proliferation and
improve other cell functions through photobiomodulation. The biphasic dose response of cells
exposed to photobiomodulating light dictates that the illumination to be evenly distributed across the
scaffold such that the cells are neither under nor over exposed to light. However, equalized
illumination has not been sufficiently addressed. Here we analyze and experimentally demonstrate
spatially equalizing illumination by three methods, namely, engineered surface scattering, reflection
by a gold mirror, and traveling-waves in a ring mesh. Our results show that nearly equalized
illumination is achievable by controlling the light scattering-to-loss ratio. This demonstration opens
opportunities for dose-optimized photobiomodulation in tissue regeneration.
Main Body:
Photobiomodulation is the use of light to regulate biological activity in cellular or tissular level. Light at red and near-infrared (NIR)
frequencies has shown to promote cell proliferation in both in vitro and in vivo experiments [1, 2]. There are three potential mechanisms
for the increased proliferation. The first mechanism is the direct absorption of red or NIR photons by cytochrome c oxidase in the
mitochondria of the cells. This absorption increases the rate of the electron transport chain, and consequently the available adenosine
triphosphate (ATP), cellular energy, of the cell, increasing the proliferation rate. The second potential mechanism is that red light (670nm
and 633nm) decreases the viscous friction of interfacial water layers surrounding the mitochondrial nano-motor, thus speeding up the
mitochondrial production of ATP [3]. The third mechanism is that light generates low levels of reactive oxygen species (ROS) within the
cell, triggering a survival mechanism that increases the production of mitochondria. This increase in mitochondria then adds to the
available ATP within the cell, which increases proliferation.
This third mechanism is optical dose dependent. With too small a dose of light, insufficient ROS will be generated and the cell will
not increase its production of mitochondria. On the other hand, if the dose is too high, too much ROS will be generated and the cell will
enter apoptosis. As such, the dose response in cell photobiomodulation has been shown to follow a biphasic response curve [4], growing
incrementally with increased dosage and leveling at some optimal range before falling precipitously (Inset, Fig 1). This biphasic dose
response requires care in the selection of power for in vivo use. If tissue is exposed from the outside of the patient, the cells close to the
source will receive exponentially more dosage than cells farther from the source, due to Beer's Law of absorption. This optical
absorption is not straightforward function of tissue type for all tissues between the source and the target cells. It is therefore reasonable to
expect that in a complex organism only a small number of subsurface target cells will reach the exact dose required for optimal effect,
while others remain either unaffected or damaged. Addressing such optimized dosage of light delivery to cells deep inside tissue
scaffolds and ultimate to the body is the motivation of the work presented here.
Tissue scaffolds are mechanical structures with engineered morphology used to hold cells while they develop to create artificially
structured tissues. Tissue scaffolds are created from biocompatible materials such as hydrogels, poly-lactic acid (PLA), and silk fibroin
by 3D printing, molding, lithography and other manufacturing techniques [5-8]. Transparent hydrogel scaffolds have been combined
with optics for sensing, optogenetics, and photobiomodulation [9]. While the transparent scaffolds do allow light to enter the tissue more
easily, they can suffer from the same dose dependence seen in external photobiomodulation where uneven illumination of the growing
tissue causes some cells to be overexposed and others to be underexposed.
In this paper we introduce and test three illumination methods for engineering transparent tissue scaffolds with equalized power
profiles. These methods can be applied with minimal changes to the mechanical structure of the scaffold, allowing them to be added to
existing scaffold designs. The first proposed method utilizes engineered diffusion; optical diffusion is the scattering of light from surface
boundaries within a media. The light-loss inside a waveguide depends on both the internal absorption [10-12], scattering events, and any
other loss mechanisms such as evanescent coupling [13], bending losses in ring structures, and proximity to metals found in the field of
plasmonics [14-18]. Regarding the material loss leading to internal waveguide absorption, this can be in principle complex if the
waveguide consists of a multitude of materials where an effective mode index determines the modal loss [19]. Here, however, we only
consider waveguides comprised of homogeneous materials. Let us now consider such as homogeneous waveguide comprised entirely of
the biocompatible material PLA extending in the x direction into the scaffold. In this case the optical power within the waveguide can be
modeled by the differential equation, Eq. 1.
(1)
where light is only lost by absorption,
, or by scattering into the fiber surrounding scaffold,
. Solving Eq. 1 for the power P at
any point within the waveguide is given by Eq. 2.
(2)
The optical power along the length of the waveguide decays exponentially with distance in the absence of gain. We wish to design the
scattering along the length of the waveguide in order to equalize the distribution of power over the surface of the waveguide. Therefore,
we seek a scattering profile,
, as a function of distance along the x-axis that provides the least power difference between input and
end of the fiber,
, with no power left at the end,
, for maximum efficiency. An optimally flat
irradiance will therefore have a derivative of zero. If we set the derivative of the irradiance to zero we need to find a function,
, that
meets the following two criteria:
as(cid:9)(cid:9)P(L)=0(cid:9)(cid:9)s(x)(cid:9)(cid:9)s(x) (3)
(4)
When we naively attempt to match
to the inverse of the exponential function, for example
, we see that
in turn
increases the exponential decay function. In fact the only function
that can satisfy (3) and (4) for all x is
, which violates
our requirement that the power at the end of the waveguide reaches 0. Hence the optimum we can accomplish with an engineered
scattering function with distance,
, is leveling the distribution over the waveguide length (i.e. waveguide surface area), as discussed
next.
Starting with transparent PLA waveguides, we explore two processes to control the scattering in transparent PLA; in the first process
we heat PLA in water causing hydrolyzation of the polymer, increasing optical diffusion (Fig. 2) [20]. In the second process NaOH is
used as a solvent to PLA to create roughness by etching the PLA surface, which again increases optical diffusion. Both methods result in
a biocompatible PLA waveguide with increased optical scattering into the scaffold.
For the second illumination method we introduce an added biocompatible gold film at the end of the PLA waveguide (Fig. 1(c)). The
gold acts as a mirror to fold or reflect the light back into the waveguide. Within certain ranges of input power, the reflected light acts to
partially compensate for the exponential decay of the original waveguide. The power reflected from the mirror can be modeled as an
additional exponential term, except now it decays from the end of the mirror, L, back to the source at x = 0.
(5)
This leads to a two-term equation for irradiance, I, from a gold-mirror waveguide where R is the mirror reflectivity.
(6)
The additional exponential term reflecting power from the end of the waveguide acts in the same manner as reflection occurs in a
Fabry-Pérot cavity [21, 22]. This levels the illumination over the length of the waveguide (Fig. 1). The gold mirror also effectively sets
the power leaving the end of the waveguide to 0, satisfying the second goal of our optimization. The difference between a waveguide's
radial power leakage with and without a mirror is summarized in Figure 1. It confirms that with a single reflection the total scattered
power is more evenly distributed along the waveguide. This leveling effect continues as the number of reflections in a Fabry-Pérot cavity
increases. It is apparent that the leveling of the spatial distribution of the scattered light is proportional to the quality (Q) factor of the
cavity. For a sufficiently-high Q-factor the average scattering function no longer depends on the distance from the source but only on
local scattering from within the cavity such as surface roughness, bending, and internal defects.
In our first experiment we treated transparent 3D printer grade filament obtained from BluMat PLA by heating the
PLA causing water hydrolyzation of the polymer to increasing optical diffusion while wrapped in damp blotting paper in a microwave
oven with varying time intervals. We then attached the treated PLA filaments to a laser light source and measured the scattering and
absorption of the PLA filaments (Fig. 2(a)).
To identify the source of the observed increased scattering (Fig. 2), we measured the change in surface roughness of the microwaved
PLA via atomic force microscopy (AFM) and measured the change in Raman absorption spectrum of the PLA. The surface roughness
was higher in the microwaved samples when compared to the control. The Raman spectrographs also showed an upward shift at the 875
cm-1 peak, indicating that both surface roughness and hydrolysis are contributing factors to the increased scattering. In our next
(cid:9)(cid:9)s(x)=emx(cid:9)(cid:9)s(x)(cid:9)(cid:9)s(x)=0(cid:9)(cid:9)s(x)experiment we capped the end of the untreated BluMat PLA filament with several micrometers of gold deposited via electron beam
deposition. The gold deposition was confined to a cap at the end of the PLA by masking all but the last millimeter of the PLA with
Kapton tape. After deposition, we measured the optical scattering from the segments to compare them to PLA segments without gold
caps (Fig. 3). Fitting the power decay resutls to eqn (6), we extrated a reflection ratio from the mirror to be approximately R = 0.7
In a final third illumination method we build on the cavity-power-leveling effect from the one-dimensional waveguide discussed
above, and introduce a ring-resonator mesh as a means to evenly distribute light across a 2-dimensional area (Fig. 4) [24]. A long photon
lifetime of a ring resonator results in an equalized spatial illumination over its surface similar to the leaky Fabry-Pérot cavity has an even
distribution of light over its surface. The single round trip electric field, E1, becomes a function of the input field [23] , Ei, via
(7)
where the loss coefficient 𝛼𝑟 is for one round trip in the ring and
denotes the change in phase as the wave travels around the
circumference of the ring when
. The total electric field, ET, is a sum of the contribution of the electric field from each
pass around the ring.
(8)
This is a geometric series and for 𝛼 < 1, we can solve for the power at any cross section of the ring.
(9)
Assuming losses around the ring exponential decay with distance, we can restate 𝛼𝑟 as a function fractional distance along the ring.
(10)
where, as before, 𝜶 is the absorption coefficient and 𝝈 is the scattering coefficient. Averaging (10) over lengths greater than 𝜆 to allow
2
𝑒𝑗𝜃
= 1, we can state the irradiance of the ring in terms of material scattering and absorption coefficients.
As with the leaky Fabry-Pérot cavity, we wish to show that increasing the quality factor decreases spatial variation of the irradiance.
To demonstrate this, we restate the electric field offset by some angle 𝜙 around the resonator.
(11)
(12)
The irradiance as a function of angle around the ring is a function of input power, and both the absorption and scattering coefficients
leading to Eq. (13).
(13)
(cid:9)(cid:9)exp(jq)(cid:9)(cid:9)q=w2pr/cWith the aim to minimize the change in irradiance with respect to angle around the ring, towards flattening the irradiance across the
ring, we rewrite (13) to become
(14)
With the total loss coefficient 𝛼𝑇 defined as 𝛼𝑇 = 𝛼 + 𝜎. We see that as the total loss coefficient 𝛼𝑇 approaches zero, the resulting
variation of irradiance over the ring also approaches zero. As light travels around a ring resonator the contribution of scattering from any
one interval around the ring becomes minimal and the ring glows evenly, independent of the location of the source. Extending this
principle to an array of ring resonators, an entire mesh-like surface can be constructed with an equal scattering distribution provided that
coupling losses are minimized. Since the ring Q is a function of both the waveguide width and bending radius, we next explore their
influence on equalizing light distribution [24]. A larger bending radius results in a smaller bending loss and vise versa. This principal
allows larger and smaller radius and waveguide width rings to be engineered into the mesh to quantitatively vary the illumination at
specific locations within the mesh. In this manner if some cells of the scaffold require a higher dose while others require a lower dose, the
scaffold itself can be designed to accommodate the cell-dependent variations in optical dose.
To test such equalized light delivery using traveling waves in ring structures, we fabricate a set of four connected ring cavities towards
demonstrating cavity-enhanced even spatial light distribution across millimeter areas using by laser cutting PLA film (Fig. 4). The rings
are 2.5mm in diameter and formed from a 790μm width waveguide. Light is sourced into the four-ring mesh through a stem of 8.5mm
length from a red laser.
The biphasic dose response of cells to optical power in photobiomodulation highlights the necessity of illumination engineering in
designing transparent, biocompatible tissue scaffolds. Here we model and demonstrate experimentally three optical methods for leveling
the illumination in transparent tissue scaffold structures. While not exhaustive, these three methods: engineering scattering, placement of
Fabry-Pérot cavities, and placement of ring resonators; form the underlying principles behind most passive illumination structures. While
spatially equalized illumination is only one factor to consider in designing transparent tissue scaffolds, without it, photobiomodulation
will be uneven in dose and the tissue scaffold will not reach its full potential. As such engineering optical nanostructures enable next
generation devices for with applications in both biomedical treatments and opto-electronics [25-31] .
Figures
Fig. 1. Modeled irradiance (a) along a 10cm length of PLA without the mirror cap, circles, the contribution of the mirror cap, diamonds, and
the total irradiance, crosses, with
, R = 0.9, shows a leveling of the power distributed over the length of the waveguide in order
to reach the optimal biphasic dose response (b), along the length of the fiber, schematic (c).
(cid:9)(cid:9)a=10[m-1](cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(a) (a) (b) dose result (b) (a) (c) (aInput Light Reflected Light Scattered Light 1.75mm Gold Mirror PLA Waveguide (cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(c) (a) (d) σ (a) α (a) - α - σ (a) (e) seconds (a) (b) Fig. 2. PLA waveguide absorption (
) and scattering
) loss test. (a) PLA waveguides were attached to a mini-laser and output power
was recorded along with an image from above while the fiber was illuminated. (b) Raman shift of the waveguide prior to treatment (dotted line)
and after treatment (solid line) at the 875cm-1 peak indicate hydrolysis. PLA waveguides were treated to incur water hydrolyzation of the polymer
to increasing optical diffusion through microwave water vaporization. Two measurements of total losses (c) scattering (d) and absorption (e)
calculated from curve fitting images and measuring the output power of 11 treated PLA waveguides shows increasing scattering with
microwaving time. This light extinction control of these waveguides allows targeting the optimal dosage on the biphasic response curve of cells
in biophotomodulation treatments.
.
Fig. 3. Light distribution effect from a leaky Fabry-Perot cavity. (a) The cavity is formed on surface-treated PLA fibers where one end is
capped with a mirror (100 nm of gold). The mirror folds light back into the fiber, thus distributing the energy more evenly along the fiber. (b)
Gold mirror cap. (c) Image of the illuminated PLA fiber with out gold mirror. (d) Image of the illuminated PLA fiber with gold mirror.
(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(b) (c) (a) (d) 1.75mm Fig. 4. A four-ring mesh fabricated from transparent 180μm thick PLA film shows leveling of the distribution of power over the surface of the
structure (a). The intensity plotted with angle where the angle 0 is the left most point in each ring, shows an average 4.9dB variance in irradiance
around each ring (b), limited by the rough edges of the laser cut PLA.
References
1. E. Mester, B. Szende, and P. Gärtner, "The effect of laser beams on the growth of hair in mice," Radiobiol.
Radiother. 9, 621 (1968).
2. K. M. AlGhamdi, A. Kumar, and N. A. Moussa, "Low-level laser therapy: a useful technique for enhancing
the proliferation of various cultured cells," Lasers in Medical Science 27, 237-249 (2011).
3. A. Sommer, M. Haddad, and H. Fecht, "Light Effect on Water Viscosity: Implication for ATP Biosynthesis,"
SCIENTIFIC REPORTS 5, 12029 (2015).
4. Y. Huang, A. C. Chen, J. D. Carroll, and M. R. Hamblin, "Biphasic dose response in low level light therapy,"
Dose-response : a publication of International Hormesis Society 7, 358 (2009).
5. R. Langer and J. Vacanti, "Tissue engineering," Science 260, 920-926 (1993).
6. C. X. F. Lam, X. M. Mo, S. H. Teoh, and D. W. Hutmacher, "Scaffold development using 3D printing with a
starch-based polymer," Materials Science and Engineering C 20, 49-56 (2002).
(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)(cid:9)1 2 3 4 (a) (b) 7. G. H. Altman, F. Diaz, C. Jakuba, T. Calabro, R. L. Horan, J. Chen, H. Lu, J. Richmond, and D. L. Kaplan,
"Silk-based biomaterials," Biomaterials 24, 401-416 (2003).
8. U. Kim, J. Park, H. Joo Kim, M. Wada, and D. L. Kaplan, "Three-dimensional aqueous-derived biomaterial
scaffolds from silk fibroin," Biomaterials 26, 2775-2785 (2005).
9. M. Choi, J. W. Choi, S. Kim, S. Nizamoglu, S. K. Hahn, and S. H. Yun, "Light-guiding hydrogels for cell-
based sensing and optogenetic synthesis in vivo," Nature Photonics 7, 987-994 (2013).
10. R. F. Oulton, X. Zhang, D. A. Genov, V. J. Sorger, and D. F. P. Pile, "A hybrid plasmonic waveguide for
subwavelength confinement and long-range propagation," Nature Photonics 2, 496-500 (2008).
11. V. J. Sorger, Z. Ye, R. F. Oulton, Y. Wang, G. Bartal, X. Yin, and X. Zhang, "Experimental demonstration of
low-loss optical waveguiding at deep sub-wavelength scales," Nature communications 2, 331 (2011).
12. Z. Ma, M. H. Tahersima, S. Khan, and V. J. Sorger, "Two-Dimensional Material-Based Mode Confinement
Engineering in Electro-Optic Modulators," IEEE Journal of Selected Topics in Quantum Electronics 23, 1-8
(2017).
13. P. O'Connor and J. Tauc, "Light scattering in optical waveguides," Appl. Opt. 17, 3226 (1978).
14. Z. Ma, Z. Li, K. Liu, C. Ye, and V. J. Sorger, "Indium-Tin-Oxide for High-performance Electro-optic
Modulation," Nanophotonics 4, 198-213 (2015).
15. K. Liu, C. R. Ye, S. Khan, and V. J. Sorger, "Review and perspective on ultrafast wavelength-size electro-
optic modulators," Laser & Photonics Reviews 9, 172-194 (2015).
16. C. Ye, S. Khan, Z. R. Li, E. Simsek, and V. J. Sorger, "λ-Size ITO and Graphene-Based Electro-Optic
Modulators on SOI," IEEE Journal of Selected Topics in Quantum Electronics 20, 40-49 (2014).
17. A. Fratalocchi, C. Dodson, R. Zia, P. Genevet, E. Verhagen, H. Altug, and V. Sorger, "Nano-optics gets
practical," NATURE NANOTECHNOLOGY 10, 11-15 (2015).
18. C. Huang, R. J. Lamond, S. K. Pickus, Z. R. Li, and V. J. Sorger, "A Sub- λ-Size Modulator Beyond the
Efficiency-Loss Limit," IEEE Photonics Journal 5, 2202411-2202411 (2013).
19. H. Huang, K. Liu, B. Qi, and V. J. Sorger, "Re-Analysis of Single-Mode Conditions for Silicon Rib
Waveguides at 1550 nm Wavelength," J. Lightwave Technol. 34, 3811-3817 (2016).
20. B. S. Ndazi, S. Karlsson, Polymerteknologi, Skolan för kemivetenskap (CHE), KTH, and Fiber- och
polymerteknik, "Characterization of hydrolytic degradation of polylactic acid/rice hulls composites in water at
different temperatures," Express Polymer Letters 5, 119-131 (2011).
21. V. J. Sorger, R. F. Oulton, J. Yao, G. Bartal, and X. Zhang, "Plasmonic Fabry-Pérot nanocavity," Nano letters
9, 3489 (2009).
22. K. Liu and V. Sorger, "Enhanced interaction strength for a square plasmon resonator embedded in a photonic
crystal nanobeam cavity," JOURNAL OF NANOPHOTONICS 9, (2015).
23. D. G. Rabus, Integrated Ring Resonators: The Compendium (Springer Berlin Heidelberg, 2007).
24. R. T. Schermer and J. H. Cole, "Improved Bend Loss Formula Verified for Optical Fiber by Simulation and
Experiment," IEEE J. Quant. Electron. 43, 899-909 (2007).
25. C. Ye, K. Liu, R. A. Soref, and V. J. Sorger, "A compact plasmonic MOS-based 2×2 electro-optic switch,"
Nanophotonics 4, 261-268 (2015).
26. K. Liu, N. Li, D. K. Sadana, and V. J. Sorger, "Integrated Nanocavity Plasmon Light Sources for On-Chip
Optical Interconnects," ACS Photonics 3, 233-242 (2016).
27. N. Li, K. Liu, V. Sorger, and D. Sadana, "Monolithic III-V on Silicon Plasmonic Nanolaser Structure for
Optical Interconnects," SCIENTIFIC REPORTS 5, 14067 (2015).
28. S. V. Boriskina, M. A. Green, K. Catchpole, E. Yablonovitch, M. C. Beard, Y. Okada, S. Lany, T. Gershon,
A. Zakutayev, M. H. Tahersima, V. J. Sorger, M. J. Naughton, K. Kempa, M. Dagenais, Y. Yao, L. Xu, X.
Sheng, N. D. Bronstein, J. A. Rogers, A. P. Alivisatos, R. G. Nuzzo, J. M. Gordon, D. M. Wu, M. D. Wisser, A.
Salleo, J. Dionne, P. Bermel, J. Greffet, I. Celanovic, M. Soljacic, A. Manor, C. Rotschild, A. Raman, L. Zhu, S.
Fan, and G. Chen, "Roadmap on optical energy conversion," Journal of Optics 18, 073004 (2016).
29. S. Sun, A. H. A. Badawy, V. Narayana, T. El-Ghazawi, and V. J. Sorger, "The Case for Hybrid Photonic
Plasmonic Interconnects (HyPPIs): Low-Latency Energy-and-Area-Efficient On-Chip Interconnects," IEEE
Photonics Journal 7, 1-14 (2015).
30. K. Liu and V. Sorger, "Electrically-driven carbon nanotube-based plasmonic laser on silicon," OPTICAL
MATERIALS EXPRESS 5, 1910-1919 (2015).
31. K. Liu, C. Zhang, S. Mu, S. Wang, and V. J. Sorger, "Two-dimensional design and analysis of trench-coupler
based Silicon Mach-Zehnder thermo-optic switch," Optics express 24, 15845 (2016).
|
1502.06416 | 1 | 1502 | 2015-02-23T13:25:52 | Differential growth of wrinkled biofilms | [
"physics.bio-ph",
"cond-mat.soft",
"nlin.AO",
"nlin.CG",
"q-bio.TO"
] | Biofilms are antibiotic-resistant bacterial aggregates that grow on moist surfaces and can trigger hospital-acquired infections. They provide a classical example in biology where the dynamics of cellular communities may be observed and studied. Gene expression regulates cell division and differentiation, which affect the biofilm architecture. Mechanical and chemical processes shape the resulting structure. We gain insight into the interplay between cellular and mechanical processes during biofilm development on air-agar interfaces by means of a hybrid model. Cellular behavior is governed by stochastic rules informed by a cascade of concentration fields for nutrients, waste and autoinducers. Cellular differentiation and death alter the structure and the mechanical properties of the biofilm, which is deformed according to Foppl-Von Karman equations informed by cellular processes and the interaction with the substratum. Stiffness gradients due to growth and swelling produce wrinkle branching. We are able to reproduce wrinkled structures often formed by biofilms on air-agar interfaces, as well as spatial distributions of differentiated cells commonly observed with B. subtilis. | physics.bio-ph | physics | Differential growth of wrinkled biofilms
Centro Nacional de Biotecnolog´ıa, CSIC, Madrid 28049, Spain
D. R. Espeso
A. Carpio∗
Departamento de Matematica Aplicada, Universidad Complutense 28040, Madrid, Spain
Center for Complex and Nonlinear Science, UC Santa Barbara, California 93106, USA
(Dated: Feb 2015, published in Physical Review E 91, 022710 (2015))
B. Einarsson
Biofilms are antibiotic-resistant bacterial aggregates that grow on moist surfaces and can trigger
hospital-acquired infections. They provide a classical example in biology where the dynamics of
cellular communities may be observed and studied. Gene expression regulates cell division and
differentiation, which affect the biofilm architecture. Mechanical and chemical processes shape the
resulting structure. We gain insight into the interplay between cellular and mechanical processes
during biofilm development on air-agar interfaces by means of a hybrid model. Cellular behavior is
governed by stochastic rules informed by a cascade of concentration fields for nutrients, waste and
autoinducers. Cellular differentiation and death alter the structure and the mechanical properties
of the biofilm, which is deformed according to Foppl-Von K´arm´an equations informed by cellular
processes and the interaction with the substratum. Stiffness gradients due to growth and swelling
produce wrinkle branching. We are able to reproduce wrinkled structures often formed by biofilms
on air-agar interfaces, as well as spatial distributions of differentiated cells commonly observed with
B. subtilis.
PACS numbers: 87.10.Mn, 87.18.Hf, 87.18.Fx
I.
INTRODUCTION
From bacterial communities to multicellular tissues,
three dimensional biological structures emerge through
poorly understood interactions between mechanical
forces and cellular processes. Being apparently less com-
plex than multicellular organisms, bacterial biofilms may
provide model systems for analyzing the interplay be-
tween cellular and mechanical features of three dimen-
sional self-organization during growth.
Biofilms are bacterial aggregates that form on moist
surfaces and are sheltered from external aggressions
by a self-produced extracellular matrix (ECM) made
of exopolymeric substances (EPS) [1, 2]. This matrix
makes them uncommonly resistant to antibiotics, disin-
fectants, flows and other chemical or mechanical agents
[3]. Biofilms are involved in chronic infections associ-
ated to biomedical implants [4, 5] and in many industrial
problems, such as biocorrosion, biofouling, efficiency re-
duction in heat exchange systems and food poisoning [6 --
8]. They are beneficial in bioremediation and biocontrol
[9, 10] or in wastewater treatments [11].
Biofilm structure depends among other variables on
the bacterial strain and the nutrient source, but is also
influenced by environmental conditions. Biofilms grown
in flows [12, 13] differ noticeably from biofilms expand-
ing on air-solid or air-liquid interfaces not exposed to
a shear force [14, 15]. These vary depending on the sur-
∗ Corresponding author: [email protected]
face, as it is the case for Bacillus subtilis forming pellicles
on air-fluid interfaces [16] or wrinkled films on agar [17].
Pattern formation is crucial to the development of biolog-
ical systems. In cellular organisms, growth is a complex
process implicating biochemical and physical mechanisms
occurring at a variety of length and timescales [18]. On
one side, genetic programs govern cellular processes, such
as growth and differentiation. On the other, mechan-
ical properties and physical forces induce macroscopic
movements of cell populations and chemical compounds
[17, 19]. Detailed models trying to describe biofilm evolu-
tion on surfaces should take into account the interaction
of these processes.
Many models have been proposed for biofilm expan-
sion in flows [20, 21], focussing typically on growth due
to nutrient consumption. The standard geometry con-
sists of a biofilm attached to a solid substratum sub-
merged in a fluid. Growth is limited by diffusion of
oxygen and nutrients from the surrounding flow into the
biofilm. EPS production favors vertical spread, pushing
cells into the flow to improve access to nutrients and oxy-
gen [1, 22]. Depending on the nutrient and oxygen avail-
ability, the bacterial strain and the hydrodynamic condi-
tions, biofilms evolve to form a variety of patterns: mush-
rooms, streamers, mounds, ripples and patchy shapes are
observed [12, 13, 23, 24]. The volume fraction of EPS in
these biofilms is large [13]. Some models view them as
cells embedded in a fluid EPS matrix [25]. Experimen-
tal observations and measurements suggest that biofilms
behave as viscoelastic materials, i.e., exhibit elastic solid-
like response to short time scale stimuli and viscous fluid-
like response to long time scale stimuli [26]. Two-phase
5
1
0
2
b
e
F
3
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
6
1
4
6
0
.
2
0
5
1
:
v
i
X
r
a
flow models of biofilm spread in fluids [27] coexist with
models that consider the growing biofilm an hyperelas-
tic material [28]. The interplay of biofilm filaments with
the flow, however, is described as interaction between an
elastic structure and a fluid [29, 30]. These models rep-
resent the microbial community as a continuum. To in-
vestigate the influence of interactions at bacterium level,
'agent' based models have been developed. Bacteria are
described as individuals whose behavior responds to ex-
ternal continuum fields. Different modeling frameworks
characterize individual bacteria in different ways. Cellu-
lar automata [31 -- 33], cellular Potts [34] and individual-
based [22, 25, 35] representations have been proposed for
biofilms in flows. We will comment on the general ap-
proaches in Section II. Basic mechanisms incorporated
in them include biofilm expansion due to cell division,
phenomenological descriptions of EPS production and
biofilm erosion caused by the external flow.
Biofilms grown in other environments show a differ-
ent structure and undergo different processes. Under-
standing their evolution requires different models, that
must take into account the specific behavior of the se-
lected bacterial strain. We consider here the develop-
ment of B. subtilis biofilms on air-agar interfaces, for
which detailed experimental evidence is becoming avail-
able [14, 15, 17, 36, 37]. Recent observations reveal the
role of cellular behavior and inner mechanical processes
on the biofilm constitution. These biofilms extract nu-
trients and water from the agar substratum itself. Cells
are densely packed, glued together by a small volume
fraction of extracellular substances [36]. The specific
autoinducer sequence triggering EPS secretion has been
identified [14]. EPS production favors water absorption
and horizontal spread [36]. Localized cell death leads to
mechanical stress relief through buckling [17].
Instead
of networks of vertical mushrooms or filaments undulat-
ing with the external flow, wrinkled films are observed
[15, 17, 37]. Partial aspects of this picture have been
modeled. Ref.
[36] analyzes accelerated horizontal ex-
pansion due to EPS production and swelling by means
of thin film approximations calibrated to experiments.
A reaction-diffusion model for the density of living cells
combined with equations for the accumulation of EPS
and waste provides understanding of the cell death pat-
terns in Ref. [17]. We introduce here a hybrid framework
that might be suitable to simulate the interplay between
the mechanical and cellular processes identified so far.
We couple a stochastic treatment of individual cellular
activities with a continuum description of elastic defor-
mations. To facilitate the coupling of stochastic cell di-
vision, death, differentiation and swelling processes with
macroscopic elastic deformations and the description of
the complex interaction with the substratum, we have
represented bacteria as cellular automata. Cells respond
to a number of concentration fields (nutrients, waste,
secreted substances), and are affected by water migra-
tion processes and elastic deformations resulting of inner
stresses due to growth, swelling and the interaction with
2
the substratum. Our numerical simulations suggest that
the wrinkle branching observed in experiments may be
the result of elastic deformations triggered by stiffness
gradients.
FIG. 1. (Color online) (a) B. subtilis biofilm grown on a 1.5%
agar surface in a biofilm-inducing medium. Photograph taken
after four days. Reprinted with the permission of Cambridge
University Press from Chai L, Vlamakis H, Kolter R, MRS
Bulletin, 36: 374-379 (2011), [14]. (b) In silico film showing
a corona of radial wrinkles emerging from a central core.
Typical wild type B. subtilis biofilms on agar show a
central core with highly connected wrinkles surrounded
by a corona with radial branches, as in Figure 1. Succes-
sive branching is also reported, see Figure 2. These struc-
tures differ from those commonly observed in biofilm pel-
licles formed on air-liquid interfaces [16]. Such pellicles
spread on a space confined by the walls of the vessel con-
taining the liquid. Ref. [16] assumes an expression for the
compression stresses and performs a stability analysis of
the equilibria of a plate model, concluding that a buckling
instability should appear above a threshold. The pres-
ence of walls stresses the pellicles as they grow, produc-
ing wrinkles. In contrast, the biofilms studied here spread
on an air-agar interface with no boundaries around. The
mechanism for wrinkle formation is bound to be differ-
3
of springs. On the other hand, neither energy nor linear
stability approaches answer the question of how different
patterns emerge and evolve from one to another. Here
we aim to compute the development of wrinkled struc-
tures as the biofilm expands horizontally, starting from
an almost flat initial state. We will solve time depen-
dent equations coupling the film to the substratum. The
residual stresses will be computed numerically averaging
information from growth and water absorption processes.
This allows to observe how wrinkles develop and how
they respond to localized variations in the microstruc-
ture. We recall below some experimental observations to
bear in mind in the sequel.
During biofilm expansion, B. subtilis populations dif-
ferentiate in several types in response to local environ-
ments created by growth, waste production, nutrient con-
sumption and cell-cell communication [14]. Cells produce
a signaling molecule, ComX. When a threshold concen-
tration is reached, a percentage of the population of cells
expresses some genes that induce a change in their pheno-
type, becoming surfactin producers. Surfactin may signal
other cells (that do not make surfactin) to produce ex-
tracellular matrix. As the biofilm thickens, starving cells
in the top become inert spores.
Initially, part of the cells are motile and express flag-
ella allowing to swarm on the surface. As the production
of ECM increases, most cells lose their individual motil-
ity [14]. Flagella are downregulated and biofilm spread
is influenced by ECM production [36].
In early stages
of the biofilm evolution, dividing cells interact with their
neighbors and push them. The biofilm becomes thicker
and the edges spread slowly on the surface. Later in-
crease in the exopolysaccharide concentration raises the
osmotic pressure, causing swelling of the biofilm due to
water intake from the agar gel [36]. Cells are pushed
from the bulk outwards, speeding up horizontal growth
and increasing nutrient intake from the agar surface.
The origin of the wrinkles seems to be controlled by
the mechanical properties of the biofilm, which are in
part governed by the production of extracellular matrix.
In early stages of the evolution of the central biofilm
core, localized cell death takes place at high density re-
gions due to biochemical stress. Cell death is usually
followed by cell shrinkage and resorption. Lateral me-
chanical stresses built up during growth are relieved at
those sites through vertical biofilm buckling, leading to
wrinkle formation [17]. Wrinkles are filled with water
extracted from agar, as in Figure 3, forming a network
of channels that enhances nutrient and waste transport
driven by surface evaporation [37]. The emergence of the
outer radial wrinkles shown in Fig. 1(a) or successive
wrinkle branching in Fig. 2(a) is less understood.
Recent developments in the study of the wrinkling of
gels and thin films on viscoelastic substrata might have
a counterpart in biofilm growth. A difference with cel-
lular films though, is that there is no feedback between
mechanical stresses, rates of swelling and growth and cel-
lular differentiation. Swelling of the outer biofilm corona
FIG. 2. (Color online) (a) B. subtilis biofilm grown on the
air-solid surface of agar gel containing water and nutrients.
Reprinted with the permission of Cambridge University Press
from Wilking JN, Angelini TE, Seminara A, Brenner MP,
Weitz DA, MRS Bulletin, 36: 385-391 (2011) [15].
(b) In
silico film showing successive wrinkle branching.
ent, even if elastic deformations are involved too. Ref.
[38] uses a neo-Hookean energy to study the elastic de-
formations of a growing circular film moderately adhered
to a rigid substrate and subject to anisotropic growth.
Tracking bifurcations in families of analytical solutions
as the anisotropy degree increases, they predict the ap-
pearance of contour undulations, delamination, buckling
on the scale of the sample after delamination and, finally,
a periodic distribution of folds. As Ref. [38] points out,
such buckling patterns are not experimentally observed
in B. subtilis biofilms. However, the analysis may be use-
ful to understand delamination and folding phenomena.
Those models do not consider the microscopic processes
occurring in a biofilm. The complex interaction with the
substratum is reduced to attachment using a collection
might produce radial wrinkles provided the center of the
biofilm was stiffer and swelled less than the corona. This
"corona instability" is commonly observed in swelling
gels when they are radially graded [39]. Such gels do
not grow but swell by absorbing water while their elastic
modulus diminishes. Contact of a film with a viscoelas-
tic substratum may also induce wrinkles, and select the
dynamics that form them [40, 41]. Wrinkle patterns that
switch smoothly from an intricate core to branching ra-
dial wrinkles have been observed in thin films on a vis-
coelastic substratum due to solvent diffusion in the film,
which creates residual stresses and stiffness gradients [42].
Based on the previous observations we propose a hy-
brid biofilm description, combining a stochastic treat-
ment of cellular processes coupled to the continuum con-
centration fields and a continuum description of defor-
mation mechanisms informed by cellular activities and
the interaction with the substratum. Biofilm growth and
physiology rely on the diffusion and transport of nutri-
ents, waste, and signaling molecules, facilitated by wa-
ter. The elastic deformation of the biofilm is described
by Foppl-Von K´arm´an equations with residual stresses
generated by growth and swelling [43 -- 45], formulated for
a thin biofilm expanding on a substratum [40, 42] con-
taining water and nutrients. We introduce a strategy to
estimate residual stresses, as well as spatial variations
in the elastic parameters of the biofilm, averaging and
smoothing stochastic information on cell type distribu-
tion, cell division and water absorption. We are able
to reproduce qualitatively patterns and trends observed
in experiments, gaining insight on the way biofilms are
shaped. We argue that stiffness gradients due to growth
and swelling may cause wrinkle branching and wrinkled
coronae. We identify governing parameters and study
their influence on the wrinkled structure. The substra-
tum properties should be carefully measured since they
largely affect the time scales for biofilm deformation.
Variations in the values of its viscosity and Poisson ratio
may change the wrinkling time scales from minutes to
days, as argued in Section IV A. Experimental observa-
tions show time variations too, compare Figures 1(a) in
Refs. [14] and [37] after four and one days, respectively.
The influence of the agar substratum on the dynamics
of biofilms and bacterial communities has already been
noticed in different contexts. Ref.
[46] shows the rele-
vance of the agarose concentration in the transition from
planar to three dimensional behavior during the invasion
of a three dimensional agar matrix by E. coli. Ref. [47]
observes that the water content of the agar substratum
affects the motility of Pseudomonas strains.
Section II describes the basic structure of the hybrid
model. Cellular processes are considered in Section III.
Stochastic division, death, differentiation and EPS pro-
duction mechanisms are introduced. Section IV discusses
film deformation according to Foppl-Von K´arm´an equa-
tions and presents water absorption strategies, together
with procedures to estimate residual stresses and spatial
variations in the elastic constants. The coupling between
4
stochastic and continuous descriptions is detailed in Sec-
tion V. Simulations of biofilm evolution activating single
or a few mechanisms are presented throughout the pa-
per. Section VI explores the coupling between processes
through a few selected tests. Finally, Section VII sum-
marizes our conclusions.
II. HYBRID MODEL
To investigate the influence of processes at bacterium
level, cells (or small clusters of similar cells) are repre-
sented by discrete 'agents'. Different modeling frame-
works describe those 'agents' and their evolution in di-
verse ways. Cellular automata distribute them over the
sites of a regular lattice [48]. The 'agents' have inter-
nal state variables, that change according to a set of
rules, stochastic or deterministic. Standard cellular au-
tomata for biofilms in flows and in porous media allocate
biomass (cells and EPS), fluids and solids over the lattice
sites [32, 33]. The dynamics of the biomass in each tile
is governed by rate equations coupled to concentrations
and fluid equations. Whenever a threshold is surpassed,
a neighboring tile is occupied, pushing biomass further
in the direction of minimal mechanical resistance. Be-
low a threshold, dead occurs with a certain probability.
Biomass fragments can detach due to erosion by the flow.
Refs. [33, 49] use this approach to predict local diffusion
coefficients, permeability coefficients and Young modulus
variations taking into account the microstructure.
Cellular Potts approaches (in particular, the Glazier-
Graner-Hogeweg variant) have been mostly applied to
describe aggregation and migration of cells and tissue
growth, see Ref. [50] for a review of applications. These
models use energy based Monte-Carlo updating and are
able to account for cell volume and shape. This is impor-
tant when describing interactions strongly dependent on
cell geometry [48]. Effective (not physical) energies typ-
ically include terms representing contact between neigh-
boring cells, a volume constraint on cells, plus chemo-
taxis terms [50]. Liquids and other substances are rep-
resented as additional cells. According to Ref. [48], the
DAH (differential adhesion hypothesis) guarantees that
final shapes correspond to the minima of effective ener-
gies. This hypothesis assumes that an aggregate of cells
behaves as a mixture of immiscible fluids [48].
Individual-based models developed for biofilms in
aquatic media consider cells as particles (usually spheres)
distributed in a three dimensional space, whose volume
and mass grow due to nutrient consumption, see Refs.
[35, 51] for a survey of applications. Cell division and
death occur when specified size thresholds are reached.
Daughter cells are distributed randomly and dead cells
are removed. The EPS matrix may be represented as
a fluid continuum, particulates and/or capsular EPS.
Biomass is eroded according to a law depending on shear.
Agent growth, division, death and shrinking cause agent
movement, via shoving algorithms and a pressure field.
5
in the system is the diffusion produced by the presence
of chemical concentration gradients of the chemical com-
pounds (oxygen, nutrients, waste products and autoin-
ducers) between the biofilm and the other phases (agar,
water and air). Cellular processes inform inner mechan-
ical stresses that shape the biofilm.
We consider a B. subtilis biofilm patch (subdomain Ω3
in Figure 3(a)) attached to an agar-gel surface (subdo-
main Ω1) and exposed to the atmosphere (subdomain
Ω2). Eventually, the biofilm may contain channels car-
rying water (subdomain Ω4) dragged from the agar sub-
stratum, as in Figure 3(b). To perform their metabolic
activities, the biofilm obtains nutrients and other sub-
stances from the gel-water substratum, and from the at-
mosphere. All the subdomains are divided in a grid of
cubic tiles. The spatial step a chosen for this description
was the size of one bacterium (around 3-5 µm) to be able
to consider individual interactions among bacteria. Each
tile can be filled with either agar, air, fluid or biofilm
biomass (either cells or ECM). Partition in tiles allows
to track the status of individual bacteria, which repre-
sent a fraction of the total biomass, and discretize the
equations governing the evolution of continuum fields.
FIG. 4. (Color online) Flow chart of the hybrid model.
The computational evolution of a biofilm in our model
is schematized in Figure 4. First, we update the cellular
status, detecting dividing, dead or differentiated cells and
allocating newborn cells and water. Then we deform the
biofilm according to its internal stresses and revise the
concentration fields and the cellular status. The next
sections describe how we take into account cellular be-
havior and mechanical processes.
FIG. 3. (Color online) Schematic representation of a biofilm.
(a) Biofilm without water network, (b) Biofilm with water
channels. For computational purposes, space is partitioned
in a grid of tiles of size a, a being the average length of a
bacterium. Each tile may be filled with air, water, agar, or
biofilm biomass (either a bacterium or ECM). This grid is
used to discretize the continuum concentration and displace-
ment fields, and to track the status of individual bacteria,
which constitute a fraction of the total biofilm biomass.
Biomass is advected following a Darcy's law involving the
pressure and the dynamic viscosity of the biofilm, which
is therefore assimilated to a fluid [35]. To investigate the
effect of the shape of bacteria on the alignment of mi-
crobial communities expanding on a surface, rod-spring
and sphere-spring representations have been recently im-
plemented [51]. Each bacterial rod or sphere is charac-
terized by its mass, velocity and position. The EPS ma-
trix is represented by springs that join the rods together,
and to the substratum. Cells divide when they reach a
threshold mass due to nutrient consumption. New cells
and springs must be created in the network. Then, the
system relaxes to a new equilibrium following Newton's
laws. A similar strategy is applied in Ref. [46] to describe
the influence of the shape of bacteria in the invasion of a
three dimensional agar matrix by E. coli.
We are interested here in understanding the elastic de-
formations caused by cell division, death, EPS produc-
tion, swelling, and the interaction with the substratum.
A framework facilitating the coupling of the relevant mi-
croscopic and macroscopic phenomena is advisable. We
resort to a cellular automata based hybrid description.
Hybrid models combine stochastic representations of cel-
lular processes with continuous descriptions of other rele-
vant fields, in our case, concentrations and deformations.
This allows to take into account the inherent randomness
of individual bacterial behaviors observed in the exper-
iments [14, 52]. It also permits to consider local inter-
actions around each cell to describe differentiation pro-
cesses.
In this model, cells are regarded as creatures living in
a grid, that reproduce, differentiate or die with certain
probability, informed by the status of the concentration
fields nearby. In a biofilm, cells are immersed in extracel-
lular matrix. The dominating mass transport mechanism
III. DESCRIPTION OF CELLULAR
PROCESSES
It is commonly accepted that biofilms are able to gen-
erate molecular signals as cell-cell communication mech-
anisms to activate gene expressions that unleash differ-
ent survival strategies adapted to the environmental con-
ditions [12, 14]. The basic approach relies on the idea
that the same organism that is sensible to certain chem-
ical inducers is also responsible of producing those com-
pounds, creating a self-feeding loop which regulates both
the genetic expression and the generation of autoinducer.
Quorum sensing mechanisms provide a way for bacteria
to count their population prior to unleashing different
behaviors such as differentiation, virulence or spreading
[53 -- 56].
Differentiation is a mechanism through which some
bacteria inside a population display physiological alter-
ations, developing different phenotypes without a change
in their genotype. These modifications lead to the de-
velopment of specialized cells with certain characteristics
which perform specific tasks.
6
and generate ComX. They are sensible to all autoinduc-
ers and may be deactivated due to a lack of nutrients or
oxygen. They may turn into surfactin generators or cells
expressing matrix genes when threshold concentrations
of ComX or surfactin are reached.
• Surfactin producers. They generate surfactin and do
not reproduce. Proliferation of these cells may be inhib-
ited by the presence of EPS producers in their neighbor-
hood.
• EPS matrix producers. They secrete EPS matrix, and
are able to reproduce at lower rate than normal cells
because their metabolism also spends resources on gen-
erating EPS.
• Inert cells, or spores. These cells do not perform any
metabolic activities, neither reproduction, nor differen-
tiation. EPS producers become spores when there is a
severe lack of nutrients in the local environment of the
cell. Spores may activate again when the environmental
conditions improve.
As time progresses, a fraction of cells expresses surfactin
genes. Later on, an increasing number of cells secretes
EPS matrix. Cells may also die as a result of biochem-
ical stresses [17]. This has been shown to occur at the
interface agar-biofilm as a result of biochemical stresses
at localized high density regions. Eventually, sporulation
starts and few normal cells remain. Spores are usually
located in upper regions of the biofilm, whereas normal
cells are restricted to the bottom edges [14].
We describe below the rules for cell differentiation,
death and reproduction.
A. Cellular division, deactivation and death
When there is an excess of oxygen, the concentration
of nutrients cn becomes the limiting concentration that
restricts biofilm growth inside the biofilm Ω3:
cn,t − Dn,b∆cn = −kn
cn
cn + Kn
.
(1)
Nutrient uptake is simply described by a Monod law.
Inside the agar substratum Ω1:
FIG. 5.
(Color online) Directional signaling in B. subtilis.
When a threshold concentration of ComX is reached, a per-
centage of the population of cells expresses genes that induce
a change in their phenotype, becoming surfactin producers.
Surfactin may signal other cells to produce extracellular ma-
trix. Starving cells become inert spores.
cn,t − Dn,a∆cn = 0.
In the water phase Ω4 (if present):
cn,t − Dn,l∆cn + v · ∇cn = 0.
(2)
(3)
Differentiation processes have been observed to depend
on the local conditions around each cell, in particular, the
state of the neighbours and the local chemical composi-
tion found around them [52, 57]. B. subtilis provides a
well known example, where initial biofilm cells may differ-
entiate in four well defined cell types. Two autoinducers,
ComX and surfactin, relate the four types of differenti-
ated cells [14], see Figure 5. The different cell types are:
• Normal cells. They perform all metabolic activities
Here, Dn,b, Dn,a and Dn,l are diffusion constants in the
biofilm, agar and water, respectively, Kn is the Monod
half-saturation coefficient, and kn(x) is the uptake rate of
the nutrient at the particular location. It will be set equal
to kn in tiles occupied by alive cells and equal to zero
elsewhere. v is the transport velocity in the water chan-
nels. On the interfaces agar-biofilm and water-biofilm
(if present) we impose transmission boundary conditions
(continuity of concentrations and fluxes). On the inter-
face biofilm-air, the interface agar-air and the edges of
the agar substratum, we have zero flux conditions. Ini-
tially, the nutrient concentration is usually taken to be
constant cn = C in agar.
Tiles occupied by alive cells C are assumed to divide
with probability [31]:
Pd(C) =
cn(C)
,
cn(C) + Kn
cn being the limiting concentration.
Initially, we real-
locate newborn cells inside biofilm by pushing existing
cells in the direction of minimum mechanical resistance,
that is, the shortest distance to the biofilm-air interface
[31, 33].
(4)
Figure 6 illustrates the expansion of a biofilm seed.
The initial round cluster grows in size in a radial way.
Brick-like colonies are formed. Their border may advance
in a wavy way depending on the velocity, as usual when a
free boundary expands outwards [58]. This spread mech-
anism may change as the biofilm develops and water is
absorbed, as explained later.
FIG. 6. (Color online) Expansion of a circular biofilm seed
driven only by nutrient consumption. The initial seed is
formed by two layers of cells with a diameter of 40 cells.
Only division and spread mechanisms are involved. Snapshots
taken every 100 steps. The initial nutrient concentration in
the substratum is C = 3Kn and the ratio kna2
DnKn
= 8.
The constants Kn affecting the probability laws are
numeric constants that should be calibrated with experi-
mental data to have a real description in the simulations
and do not have to match with saturation constants used
in the source terms of the reaction-diffusion equations for
the concentrations, which are related with a chemical sat-
uration limit. To simplify, they have been chosen to agree
with the values for saturation constants.
7
Cells may suffer a shortage of nutrients, triggering a
sporulation mechanism to preserve bacteria in a "deacti-
vated" state until the environmental conditions improve.
To reflect this, cells with a local limiting concentration
below a certain threshold will become spores in our simu-
lations. We deactivate cells C with probability 1 − Pd(C)
whenever Pd(C) < γd for a certain threshold γd > 0.
These cells may activate again when the conditions im-
prove.
As mentioned before, cells may also die due to bio-
chemical stress in high density regions where waste prod-
ucts and toxins accumulate but nutrients deplete. Taking
for instance the concentration of waste cw at a location
as an indicator of death, a cell C is scheduled to die with
probability:
Pw(C) =
cw(C)
cw(C) + Kw
,
(5)
whenever Pw(C) > γw for a certain threshold γw > 0.
The concentration of waste is governed by
cw,t − Dw,b∆cw = kw
δi,w,
(6)
i
inside the biofilm Ω3. δi,w(x) equals 1 in regions occupied
by alive cells Ci and vanishes otherwise. On agar-biofilm
and air-biofilm interfaces we impose no flux boundary
conditions.
If a fluid phase Ω4 is present, we impose
transmission boundary conditions at the interface and
include a diffusion-convection equation similar to Eq. (3)
for waste transport in the liquid. Dw,b, Dw,l are the
diffusion constants, and kw the production rate of waste
and toxins at the particular location.
B. Surfactin and EPS production
(cid:88)
(cid:18)
(cid:19)(cid:88)
Normal cells may turn into surfactin generators when
a threshold concentration of ComX is reached. ComX
is a molecule related with quorum sensing mechanisms
in B. subtilis [14]. It is produced by all the cells in the
biofilm, except those which are not metabolically active.
The concentration ccx of ComX inside the biofilm Ω3 is
governed by:
ccx,t − Dcx,b∆ccx = kcx
1 −
ccx
ccx + Kcx
i
δi,cx, (7)
with no flux boundary condition on agar-biofilm and air-
biofilm interfaces. kcx is the production rate, and Dcx,b
the diffusivity. δi,cx(x) equals 1 if cell Ci produces ComX
and vanishes otherwise. kcx is multiplied by an inhibition
factor for large concentrations. A guess for Kcx may be
the half-saturation constant of ComX. If a liquid phase
Ω4 is present we impose transmission boundary condi-
tions at the interface and include a convection-diffusion
equation similar to Eq. (3) for ComX transport in the
liquid. Normal cells C are assumed to become surfactin
producers with probability:
8
Ps(C) =
ccx(C)
ccx(C) + Kcx
(8)
whenever Ps(C) > γs. The threshold γs reflects that dif-
ferentiation should start when a minimum background
concentration of ComX is reached.
,
ccx
ccx+Kcx
When EPS producers are present, they act as in-
hibitors. We may set Ps = Pinhib
with Pinhib
the ratio of neighbors producing EPS to total number
of neighbors. If a cell is surrounded by EPS producers,
there is no any chance of differentiation into a surfactin
generator. A similar effect is obtained setting a much
lower threshold for differentiation into a surfactin pro-
ducer compared to EPS producers, as we usually do in
our simulations.
Surfactin acts as an autoinducer in normal bacteria,
changing their phenotype to become EPS producers [36].
The concentration cs of surfactin inside the biofilm is
governed by:
cs,t − Ds,b∆cs = ks
1 −
cs
cs + Ks
δi,s,
(9)
(cid:18)
(cid:19)(cid:88)
i
with no flux boundary condition on agar-biofilm and air-
biofilm interfaces. ks is the production rate, Ds,b the dif-
fusivity and Ks the half-saturation of surfactin. δi,s(x)
equals 1 if cell Ci produces surfactin and vanishes oth-
erwise. When a liquid phase Ω4 is present transmis-
sion boundary conditions hold at the interface and a
convection-diffusion equation similar to Eq. (3) governs
surfactin transport in the liquid. We assume that ac-
tive cells C not secreting surfactin become EPS producers
with probability:
(cid:19)(cid:18) cs(C)
(cid:19)
cs(C) + Ks
,
(10)
(cid:18)
Pe(C) =
1 −
cn(C)
cn(C) + Kn
cs(C)
cs(C)+Ks
> γe1 , which means that a minimum
whenever
concentration of surfactin is required. The probability
increases with the availability of surfactin and the scarce-
ness of nutrients. This is consistent with measurements
reported in Ref.
If we want to enforce a level of
nutrient depletion to allow differentiation into EPS pro-
ducers we may further impose
< γe2. This may
forbid differentiation below a certain height.
cn(C)+Kn
cn(C)
[59].
Figure 7 illustrates the distribution of cells when we ac-
tivate the differentiation processes, in combination with
standard growth and spread. As the size of the biofilm
increases, the concentration of ComX raises until the
threshold γs is reached. At this point bacteria start to
differentiate and secrete surfactin. This triggers differ-
entiation of cells into EPS producers when the surfactin
threshold γe1 is reached. As the number of EPS cells
increases, the rate of differentiation into surfactin gen-
erators stabilizes and may even decay due to an inhibi-
tion mechanism induced by the EPS cells [14]. Relative
proportions for each type of cell were adjusted by fitting
FIG. 7. (Color online) Slices of a biofilm showing the cell type
distribution during early stages of the differentiation process.
All the slices correspond to the same biofilm and are taken
at the same time. The evolution started from an initial seed
with a diameter of 60 cells and variable thickness, represented
by three small peaks. Dead cells appear first where the initial
peaks where located. In later stages, inert cells proliferate at
the top whereas normal cells are restricted to the bottom and
the edges. Each colored box is one cell. Parameter values:
= 0.01,
C = Kn,
ksa2
DsKs
0.00001,γe2 = 0.2.
= 0.8, γw = 0.005, γd = 0.00001, γs = 0.001, γe1 =
kcxa2
DcxKcx
kw a2
Dw Kw
kna2
DnKn
= 0.001,
= 0.01,
parameters, but there should be an experimental calibra-
tion to improve these values. Normal cells are found in
the edges and near the interface biofilm-agar. EPS pro-
ducers appear once surfactin is generated and a certain
thickness is reached so that the nutrient concentration
depletes. Dead cells concentrate on the bottom surface
at inhomogeneity peaks when the initial seed has variable
height, due to higher waste concentration. As the biofilm
increases its thickness, scarcity of nutrients is felt in the
upper part due to the high resistance to mass transfer
across the increasing thickness. This effect induces the
appearance of spores in the upper part, slowing down the
growth of the biofilm. The geometric distribution of each
type of cell seems to follow qualitatively trends reported
in Refs. [14, 17, 59].
Implementing the differentiation strategy described
above requires measuring a number of kinetic constants
for different concentrations. Since these values are cur-
rently unknown, we have scaled the parameters to obtain
the expected behaviors in small clusters of cells, lower-
ing the computational cost. For Gram negative Vibrio
harveyi, quorum sensing processes triggering biolumines-
cence have been modeled in terms of feedback loops [56]
and applying information theory to the quorum sensing
circuit [60], which may reduce the number of parameters
involved. Signaling parameters have been deduced from
in vivo data for V. harveyi [61].
For simplicity, the biofilm in Figure 7 does not include
the ECM phase, only cells. We might include it following
the procedure described below. Once EPS production
starts, a new concentration ce diffuses in the environ-
ment:
ce,t − De,b∆ce = ke
1 −
ce
ce + Ke
δi,e.
(11)
(cid:18)
(cid:19)(cid:88)
i
ke is the production rate, De,b the diffusivity and Ke the
half-saturation of EPS. δi,e(x) equals 1 if cell Ci is an EPS
producer and vanishes otherwise.
Initially, these sub-
stances are just compounds diffusing in the environment.
As the concentration increases they become a polymeric
envelop for the cells. This description can be improved
taking into account the presence of molecules of different
size. Monomers diffuse, but longer molecules stay where
they are produced. The concentration of momoners is
governed by a diffusion equation coupled to the concen-
tration of larger molecules, whereas the latter is governed
by rate equations, that substitute Eq. (11).
Images in Ref.
The volume fraction of biofilm occupied by extracel-
lular substances varies largely depending on the type
of biofilm.
[13] show biofilms in flows,
formed by scattered cells embedded in a dominant EPS
phase. On the contrary, imaging biofilms on air-semisolid
interfaces at hight resolution shows a structure of densely
packed cells glued by EPS, see references [36], [14] and
[17]. Space is mostly occupied by cells. This suggests
skipping the creation of an EPS phase and thinking of it
as a virtual glue that keeps cells together and modifies
the mechanical properties of the biofilm, in particular, its
elastic constants and the osmotic pressure. If we choose
to create an EPS phase, we may exploit the concentra-
tion ce of EPS to generate new biofilm tiles filled with it.
At any tile, we set
Pm(C) =
ce(C)
ce(C) + Ke
.
(12)
If this value is larger that a threshold γm, we create an
EPS tile at that location and shift biomass around in the
direction of least mechanical resistance. γm should be
calibrated so that the fractions of biomass corresponding
to cells and EPS fit the experimental observations. The
true cells are a fraction of the total amount of tiles occu-
pied by biomass. They divide and differentiate according
to the same probabilistic rules we have already described.
9
Individual-based models for biofilms on agar surfaces
represent the EPS as springs joining the cells [51]. For
biofilms in flows, expressing larger fractions of EPS, each
'individual' contains cellular mass and capsular EPS,
that vary according to user defined rate equations. Ad-
ditional EPS particulates may be created as EPS pro-
duction increases following user defined rules [35]. Refs.
[22, 62] study EPS dynamics in a quorum sensing frame-
work, defining competing strains: EPS producers, EPS
non producers, strains that downregulate EPS produc-
tion for large cell densities. This is not the reported quo-
rum sensing circuit for B. subtilis, see Ref.
[14]. Cellu-
lar automata models for biofilms in flows distribute cel-
lular mass and EPS in each lattice site following phe-
nomenological rate equations [32, 33]. Spread to neigh-
boring sites occurs as a threshold mass is surpassed. In
both frameworks, cells are scheduled to die when their
mass or volume falls below a threshold. No memory of
their former presence, relevant for elastic deformations,
is kept. Neither the relevance of waste accumulation for
cell death, pointed out in Ref.
[17], nor the currently
known quorum sensing circuits for EPS production are
incorporated in those models. Dynamic energy budget
descriptions of cell dynamics [63] might provide a sophis-
phicated framework to describe the evolution of individ-
ual cells taking into account their mass, volume, mainte-
nance, substance secretion and the presence of nutrients,
oxygen and toxicants. However, extensions to biofilms
are yet to be devised.
C. Nondimensionalization of the concentration
equations
Nondimensionalization of the equations for the con-
centrations listed above reveals a key governing param-
eter for each concentration: F = ka2
, where k is the
KDb
uptake or production rate, a the bacterium size, K the
half-saturation constant and Db the diffusivity inside the
biofilm. Setting c = c
a , the equations for
concentrations become:
K and x = x
− ∆xc = F f
,
(13)
(cid:16) c
(cid:17)
c + 1
inside the biofilm, where f represents the production or
consumption terms, and
(14)
− ∆xc + v · ∇xc = 0,
in the water channels, where v = v a
. We have sup-
D(cid:96)
pressed the time dependence because diffusion of chemi-
cal species is faster than the cellular processes considered.
For each biofilm configuration, we solve these nondimen-
sionalized stationary equations with the corresponding
boundary conditions by a relaxation technique. Then
we compute all the probabilities for division, death and
differentiation, detect the different types of cells and allo-
cate newborn cells. This generates a new biofilm config-
uration for which the procedure is repeated. A complete
reproductive cycle of a bacterium was set as the basic
time step for the cellular processes in the model (which
is typically in the order of 26-45 minutes depending on
nutrient and temperature). This choice implies that our
system is discrete in space and time [31].
IV. DESCRIPTION OF MECHANICAL
PROCESSES
Differentiated cells produce chemicals, which may in-
duce further differentiation of other cells or trigger me-
chanical processes that alter the biofilm structure, im-
proving its chances of spread and survival. EPS produc-
tion increases the stiffness of the biofilm [17], which may
be identified with a material undergoing macroscopic de-
formations caused by growth [16] over a substratum. The
spatial distribution of dead and inert cells, together with
water absorption, may induce additional stiffness gradi-
ents. This motivates allowing our biofilm to move along
the tiles according to microscopically informed contin-
uum descriptions of the underlying mechanical processes.
A. Out-of-plane deformations and wrinkle
formation
Due to the interaction of components in the extracellu-
lar matrix, dissipative phenomena take place inside soft
tissues during deformation processes and the mechani-
cal response of tissues if often viscoelastic. Nevertheless,
growth induced deformations on time scales much larger
than those of relaxation may be regarded as purely elas-
tic [45]. Since the height of real biofilms is about 10-100
times smaller than their radius and they are initially thin
and flat, we may use Foppl-von K´arm´an [43] equations
for thin elastic plates to study their elastic deformation
reducing dimensionality [16]. Residual stresses due to
growth are incorporated in the description in terms of
the growth tensor, see Refs.
[44, 45]. The shape of the
body is then determined by both growth and elastic de-
formation. The equilibrium deformation of a growing tis-
sue is described by a variant of the Foppl-Von K´arm´an
equations [45]:
D(∆2ξ − ∆CM ) − h
∂
∂xβ
σα,β
(cid:18)
(cid:19)
∂ξ
∂xα
∂σα,β
∂xβ
= P,
(15)
= 0,
(16)
10
(cid:19)
(cid:18) ∂uα
∂xβ
growth term. The components of the tensor g are the
growth rates, that is, the rate of volume supply per unit
volume. Defining consistently a growth tensor g is a non
trivial issue. A possibility is to set g = ∇w [64], w being
the vector field of displacement of growing biomass in the
biofilm w, which is determined by the cellular division
and spread mechanism and water absorption processes.
Stresses σ and strains ε are defined in terms of in-plane
displacements u = (ux, uy) [39, 45] by:
εα,β =
1
2
+
∂uβ
∂xα
+
∂ξ
∂xα
∂ξ
∂xβ
+ ε0
α,β,
(17)
σxx =
E
1 − ν2 (εxx + νεyy), σxy =
E
1 + ν
εxy,
(18)
σyy =
E
1 − ν2 (εyy + νεxx).
The Poisson ratio of different tissues has been measured
to be about 0.49999 [65]. We will set the Poisson ratio of
the biofilm ν = 1
2 (incompressibility) [45]. The average
elastic modulus E of a wild type B. subtilis biofilm has
been measured to be about 25 kP a [17]. When the resid-
ual strains ε0
α,β are due to growth, they are expressed in
terms of the growth tensor as:
α,β = − 1
ε0
2
(gαβ + gβα + gzαgzβ) .
(19)
A growing biofilm is in a state of compression due to cell
division and, eventually, water absorption. Alternatively,
this may be represented by a residual strain
α,β(x, t) = −ε0(x, t)δα,β,
ε0
ε0 > 0.
(20)
If we assume that cells do not grow at expense of their
neighbors, δα,β is a diagonal unit tensor in polar coordi-
nates in a circular film.
For a biofilm growing on a surface no external loads act
on the edges, therefore, P = 0. However, the interaction
with the surface affects the deformation process and has
to be included in the description. A correction to Eqs.
(15)-(16) applicable to thin elastic films growing on a
viscoelastic substratum is proposed in Ref. [40]:
(cid:34)
D(−∆2ξ + ∆CM )
(cid:19)(cid:35)
∂ξ
∂t
=
1 − 2νv
2(1 − νv)
hv
ηv
(cid:18)
+ h
∂
∂xβ
σα,β(u)
∂ξ
∂xα
∇ · σ(u) − µv
ηv
u,
∂u
∂t
=
hvh
ηv
− µv
ηv
ξ,
(21)
(22)
where α, β stand for x, y and summation over repeated
indexes is intended. The first equation describes out-of-
plane bending ξ(x, y) and the second one in-plane stretch-
ing for the displacements u = (ux(x, y), uy(x, y)). (x, y)
vary along the 2D projection of the 3D biofilm structure.
The bending stiffness is D = Eh3
12(1−ν2) , and h the initial
plate height. P is the external pressure at the border of
the sample. CM = ∂(gzx+gxz)
is a residual
+ ∂(gzy+gyz)
∂x
∂y
where hv is the thickness of the viscoelastic substratum
and µv, νv, ηv its rubbery modulus, Poisson ratio, and
viscosity, respectively.
The equilibrium equations describe possible equilib-
rium configurations, and yield information on their sta-
bility or changes of stability. To describe the dynam-
ics of wrinkle formation and the finally selected pattern
we must solve the time dependent equations. Equations
(21)-(22) were derived for the interface between the thin
film and the substratum. This interface moves vertically
following the displacement field ξ. This shifts all the tiles
in the three dimensional biofilm-substratum system ver-
tically.
Choosing new dimensionless variables:
11
x =
x
L
,
u =
u
L
,
ξ =
ξ
h
,
σ =
σ
E
,
t =
t
T
,
and setting L = γh, the dimensionless equations read:
(cid:34)
∂ ξ
∂t
=
12(1 − ν2)γ2 ∂
∂ xβ
+ (−∆2
x
ξ + ∆x CM )
(cid:33)
∂ ξ
∂ xα
(cid:32)
(cid:35)
σα,β(u)
− T
µv
ηv
ξ,
∂ u
∂t
= τ ∇x · σ(u) − T
µv
ηv
u,
with
T =
12(1 − ν2)γ4
2(1 − νv)
ηvh
1 − 2νv
hv
E
(1 − νv)
(1 − ν2)γ2.
(1 − 2νv)
τ = 24
(23)
(24)
(25)
,
We set L equal to the maximum averaged radius of the
biofilm in study. Common experimental values for the
parameters defining the spatial scales are h ∼ 100µm,
hv ∼ 100h. The size of the biofilm may vary with the
nutrients and the agar gel nature. Figure 1(a) in Ref.
[14] shows a wrinkled biofilm with L ∼ 50h grown in
four days on a 1.5% agar surface. Figure 1 (a) in Ref.
[37] reproduces wrinkled biofilms with L ∼ 50h within
one day.
u and T µv
ηv
The linear terms T µv
ηv
ξ are just damping
contributions from the substratum slowing down the evo-
lution. We will neglect them in our preliminary stud-
ies. Wrinkling behavior is observed when the coefficient
12(1 − ν2)γ2 is large enough, so that the contribution
of nonlinear effects and residual stresses in Eq. (23) be-
comes relevant and drives the plate out of the flat equi-
librium state, see Figure 8.
The factor τ in Eq. (24) separates the time scales for
the evolution of u and ξ. The closer to 0.5 the Poisson
ratio νv is, the larger this separation is. The in-plane dis-
placements u may reach a quasistatic situation in which
their evolution is driven by slower changes in the resid-
ual stresses created by growth, swelling, and off-plane
displacements.
The constitutive parameters for an agar gel may vary
depending on the agar source, its content of agar and
other products, in particular, water. When an hydro-
gel is fully swollen, its mechanical behavior is similar to
those of rubber-like materials [66], which have a Poisson
ratio of about 0.5. For some agar gelatines the Young
modulus is measured to be 0.4999, with an elastic mod-
ulus about 27 kPa [65]. In other agar gels these numbers
FIG. 8. (Color online) Connected network of wrinkles switch-
ing from an intricate core to radial branches, formed in a ho-
mogeneous circular film when a circular front of residual com-
pression stresses of magnitude ε0 = 0.1 expands at constant
velocity. Parameter values: ν = 0.5, E = 25 kP a, νv = 0.45,
µv = 0 and γ = 16. dx = 0.1 h, where h is the film thickness
before wrinkling.
go down to νv = 0.32 with an elastic modulus Ev = 52
kPa [67]. This modulus increases with the agar concen-
tration from ∼ 30 kPa in 0.5% samples up to 700 kPa in
5% samples. The viscosity ηv of agar gels is reported in
the range of 100 cp = 0.1 Pa·s for 1.5% samples, see Ref.
[68]. As the agarose concentration increases to 8%, the
viscosity raises to 1− 1.5 Pa·s [69]. These values increase
as the temperature diminishes. According to Ref.
[69],
no measurements can be obtained below 36o C. Standard
experiments are carried out at room temperature, usu-
ally below that threshold. Effective values for viscosities
in a MPa·s range are adjusted in Ref. [70].
A large variability in the time scales may arise due
to uncertainty in the experimental values of νv and ηv.
Setting for instance νv = 0.4999, ηv = 0.1 Pa·s, we find
T ∼ 31 hours. This value varies enormously depending
on νv, that changes with the water content. Switching to
νv = 0.45 and ηv = 1 Pa·s, we find T ∼ 2480 s ∼ 40 m.
The time scale scale for u is T (cid:48) = T
τ ranges from 10−3 s
in the first case to 10−2 s in the second one. If ηv enters
the range of kPa·s or MPa·s, this value increases by a
factor 103 − 106.
B. Water absorption and water channels
As mentioned earlier, variations in the osmotic pres-
In Ref.
sure extract water from the agar substratum.
[36], a macroscopic model for water migration from the
agar substratum into the biofilm is proposed in terms
of the volume fractions of water and biofilm at each lo-
cation. The biofilm is formed by cells and extracellular
matrix, which is often considered a gel with ability to
swell. As observed in Ref. [37], flat regions of the biofilm
are very resistant to flow. Water cannot be driven inside
without fracturing the biofilm or delaminating it from
the substratum. By contrast, water flows beneath the
wrinkles, forming an intricate network of channels [37].
Wrinkles origin in the core of the biofilm associated with
dead areas [17]. Water flows along the channels driven
12
by surface evaporation [37].
These observations suggest a strategy to incorporate
water into our biofilm description. Water absorption
by the biofilm increases the volume of the cells and the
EPS phase. Enlarging the size of the tiles to reflect that
fact is unpractical. We resort instead to inserting water
tiles in the biofilm, representing water absorption by the
biomass. In that way, the biofilm contains a volume frac-
tion of adsorbed water that changes its volume. A low
cost strategy, easy to implement in our hybrid framework
is the following. As in Ref.
[36], we propose for the os-
motic pressure a law of the form π = Πφ, but replacing
the volume fraction φ of biomass by our information of
the biomass available at biofilm columns
π = Π
N
Nmax
,
(26)
where N is the number of tiles occupied by biomass in
a column and Nmax is the total number of tiles in the
column, including water. Water tiles are created with
probability:
Pl(C) =
π(C)
,
π(C) + E(C)
(27)
where E(C) is the Young modulus at tile C and the height
of the columns in Eq. (26) is updated as we create wa-
ter tiles. The status of neighboring tiles is shifted in the
direction of minimal mechanical resistance, except when
water occupies a dead cell. More water will be absorbed
in the columns containing a larger fraction of biomass,
and in softer regions. Scattered inner tiles do not consti-
tute a separate phase, but are considered to be part of
the swollen biomass.
We compute E(C) modulating the measured reference
value of the Young modulus of the biofilm with an av-
erage of weights representing status of the neighboring
tiles. Weights range from zero for water, a small posi-
tive value for dead cells, a larger value for normal cells,
up to one for EPS producers. Alive cells attached to the
substratum contribute much larger weights than the rest.
E(C) will be small wherever we have dead cells. As Fig-
ure 9 shows, those regions fill with water in successive
steps, becoming a water phase that expands following
the wrinkles. Combining water transport through those
channels, with the cellular processes described in Section
III, we see that cell death due to waste accumulation is
reduced and the colony is able to expand in a sustained
way. In early stages we do not allow newborn cells into
this water phase due to the pressure gradients that drift
water from the agar gel into the biofilm. As the water
content of the agar gel diminishes we might allow cell
expansion around the channels, as observed in Ref. [37].
This would keep water inside the biofilm.
In Figure 9, we have distributed dead cells above the
substratum, in the wrinkled regions. The value of Π is
chosen small enough to avoid water proliferation inside
the biofilm, about E/20. These simple probability laws
produce compression residual stresses that vary with the
FIG. 9.
(Color online) (a) Biofilm-substratum system de-
formed following out-of-plane displacements ξ predicted by
Eqs. (21)-(22) for residual stresses expanding as a radial com-
pression front. (b) Bottom slice of (a) displaying the arrange-
ment of biofilm and substratum. (c) Intemediate slice of (a)
showing the allocation of dead cells. (d) and (e) Accumula-
tion of absorbed water in slice (c) triggered by the presence
of weakened dead areas after 1 and 40 steps of the water ab-
sorption process, respectively.
biofilm height, as shown in Figure 10, where Π = E/3.
Values in the range Π ∼ E are suggested in Ref. [15].
A time scale for water absorption processes may be
inferred from the analysis in Ref. [36], where a thin film
description of biofilm spread due to swelling and growth
is proposed. For radial biofilms and small departures
from osmotic balance, a similarity approximation for the
height h and the radius R of the biofilm yields a time
scale 1/q for the spread process, where q is the rate of
biomass production.
We cannot directly couple the model in Ref. [36] to a
cellular automata description of cell division, death and
differentiation to describe biofilm spread due to swelling.
The main reason is that it already includes a source rep-
resenting growth. An alternative model of fluid transport
in the biofilm would be necessary for a better description
of the water absorption process and biofilm spread in our
framework.
Once water channels have been carved in the biofilm,
we might couple our cellular automata description to the
hydrodynamics of the flow as done for biofilms in porous
media [33]. This assumes that the deformation of the
biomass and the extracellular fluid flow can be computed
in a sequential manner, which happens when the biomass
behaves as an elastic material or the flow induced de-
formation of the biomass matrix is negligible [71]. Us-
ing order-of-magnitude analysis, Ref. [72] establishes do-
mains of validity of different models for liquid transport
in a fluid-solid system depending on the volume fraction
and viscosity of the fluid, the volume fraction, elastic con-
stants and density of the solid, the hydraulic permeability
of the system, the characteristic times for displacement
of the solid, and the characteristic macroscopic length of
the system.
C. Growth tensor and elastic constants
The components of the tensor g entering Eqs. (23)-(24)
are the growth rates, that is, the rate of volume supply
per unit volume. New tiles may add to our biofilm due to
cell division, EPS creation or water adsorption. Wher-
ever they are created, they shift another tile in the direc-
tion of minimal mechanical resistance. The growth ten-
sor at each grid location (ix, iy, iz) is computed by keep-
ing track of all the new tiles inserted and the direction
in which their predecessors where shifted. We define a
vector w = (w1(ix, iy, iz), w2(ix, iy, iz), w3(ix, iy, iz))a.
w1(ix, iy, iz) is determined by cumulatively adding ±1
for each tile shifted in the x direction in the positive or
negative sense, respectively. w2 and w3 are evaluated in
a similar way, along the y and z directions. The final
vector is normalized to have norm a. Then, we com-
pute ∇w, where the derivatives are approximated by fi-
nite differences that use the known grid values. To es-
timate g(ix, iy) we consider all the contributions from
∇w(ix, iy, iz) for varying iz.
The resulting tensor usually presents abrupt oscilla-
tions due to stochasticity, constrained motion along a
restricted set of directions in a cubic grid and varying
biofilm thickness. Such oscillations in the grid scale may
destabilize numerical solutions of Eq. (23). Averaging
and smoothing these tensors to avoid numerical instabil-
ities, we obtain residual stresses that serve as a basis for
the numerical tests, see Figure 10. Averaging rapidly os-
cillating coefficients and sources is a standard practice to
study macroscopic deformations of materials.
The reference Young modulus of the biofilm is esti-
mated from experimental measurements [17]. The knowl-
edge of its microstructure gained from our simulations al-
lows to introduce spatial modulations. At each location,
we may regulate the reference value multiplying it by the
sum of weights describing the status of its neighboring
tiles, as we did in section IV B. A possible qualitative
choice for those weights are zero for air, a small fraction
13
FIG. 10. (Color online) Residual stresses ε0
xx for the biofilm
seeds depicted in (a), (b) and (c). (d), (e) and (f) represent
their values after one step of the stochastic division and spread
process, with C = 2Kn and kna2
= 0.01. Spatial variations
DnKn
related to thickness are identifiable in (g), (h), (i), obtained
averaging a few trials of the stochastic procedure. These fields
approach their average values (about −0.1). A single trial
already gives the average value used in rough approximations
by constants. (j), (k), and (l) show the effect of including dead
spots in the biofilm seed. Localized depressions are observed.
(m), (n), (o) reproduce the residual stresses generated by the
water absorption scheme when Π = E/3. The average values
are about −0.2. ε0
xy is about
3-2 times smaller.
xx in all cases. ε0
yy resembles ε0
of one for dead cells, a larger fraction for normal cells
and one for cells producing EPS. The biofilm becomes
softer where it swells easily. In this way, we obtain spa-
tially dependent elastic moduli that allow to investigate
variations in the macroscopic deformation of the biofilm
in response to changes in its microstructure.
V. COUPLING OF CELLULAR AND
MECHANICAL PROCESSES
The interaction between the previous descriptions of
cellular and mechanical processes is visualized in Figure
4. The scheme below details a practical implementation.
1. Initialization. A matrix is created to indicate the
status of the tiles in the computational grid of step
a. At first, we usually set S(ix, iy, iz) = 2 in the
regions occupied by a biofilm seed, assuming that it
is formed by undifferentiated cells stuck together.
As tiles (ix, iy, iz) are filled with differentiated cells,
EPS or water in the next steps, S will take different
positive integer values. Sections occupied by air or
substratum correspond to zero or negative values.
This grid is used to keep track of cellular processes
and also to discretize continuum equations. As the
biofilm enlarges, we might resort to a coarser grid
for the latter purpose. Initially, the concentration
of nutrients is set equal to a constant C in the sub-
stratum and zero elsewhere. The concentrations of
waste, ComX, surfactin and EPS are set equal to
zero everywhere.
2. Cellular processes.
(a) Concentration update. The concentrations of
nutrients, waste and ComX are evaluated solving
Eqs. (1)-(3), (6) and (7) with the specified bound-
ary and initial conditions. Whenever surfactin or
EPS producers are present, Eqs. (9) and (11) are
solved too. All the equations are nondimensional-
ized as indicated in expressions (13)-(14). The so-
lutions of the time dependent diffusion problems for
these chemicals relax to their stationary states in a
short time, compared to the typical time scales for
cellular processes. Explicit finite difference schemes
provide a low cost approximation to the stationary
concentrations after a number of steps. Those ap-
proximations are stored and used in the next stages
to evaluate behavioral probabilities.
(b) Probabilities for cell activities. They are only
computed at tiles (ix, iy, iz) occupied by cells. Cells
that are not already dead are killed with probabil-
ity Pw whenever the concentration of waste sur-
passes a threshold γw. We evaluate expression
(5) and generate a random number r ∈ [0, 1].
If
max(r, γw) < Pw, we kill the cell. Alive cells de-
activate with probability 1 − Pd, Pd given by for-
mula (4), as long as the nutrient concentration is
low enough. Active cells become surfactin produc-
ers with probability Ps defined in formula (8) if the
concentration of ComX surpasses a threshold. Ac-
tive cells not releasing surfactin become EPS pro-
ducers with probability Pe given by formula (10)
when a minimum surfactin level is reached and the
nutrient concentration is depleted.
14
(c) Cellular division and spread. Active cells not
secreting surfactin divide with probability Pd. The
newborn cells are placed in neighboring tiles in the
direction of minimal mechanical resistance. In early
stages, this may be taken to be the shortest dis-
tance to the biofilm-air interface.
(d) EPS phase. Accumulation of EPS may prompt
the creation of an EPS phase according to Eq. (12).
(e) Growth tensor and elastic parameters. As EPS
production increases, the biofilm acquires consis-
tency. We have used in our simulations average
measured values of the Young modulus. Informa-
tion on the spatial distribution of different types of
cells allows to spatially modulate that value. Com-
puting the out-of-plane deformations in the next
stage requires the previous derivation of a growth
tensor from the biomass growth process. This is
done as indicated in subsection IV C.
3. Mechanical processes.
(a) Water absorption and water phase. EPS pro-
duction changes the osmotic pressure and prompts
water migration from the substratum towards the
biofilm. Water absorption by the biomass may be
accounted for by means of formula (27). Dead
zones near the surface will easily fill with water,
which may expand along the wrinkles between the
biomass and the substratum. A water phase is cre-
ated. As the volume fraction of water increases and
water channels develop, a description of liquid flow
in the system might be necessary.
(b) Growth tensor and elastic parameters.
Infor-
mation on the spatial distribution of absorbed wa-
ter allows to spatially modulate the average Young
modulus, decreasing it in swollen regions. The
growth tensor is updated to reflect the volume sup-
ply due to water absorption. This is done as indi-
cated in subsection IV C.
(c) Out-of-plane displacements. Vertical displace-
ments due to internal stresses are estimated solving
Eqs. (21)-(22). The equations are nondimension-
alized following Eqs. (23)-(24). Basic explicit dif-
ference schemes provide low cost approximations.
Alternatively, faster spectral methods may be used
[40, 42]. As the biofilm becomes more heteroge-
nous we might need to solve a three dimensional
elasticity problem or include lower order terms in
the Von-K´arm´an equations.
(d) Biofilm deformation. Tiles (ix, iy, iz) at the
interface biofilm-substratum are shifted vertically
a number of tiles equal to the integer part of
ξ(ix, iy)/a, pushing their neighbors in the three di-
mensional biofilm. This is done shifting the status
of such tiles in the matrix S(ix, iy, iz). Relative
status of neighbors in the vertical direction is pre-
served.
4. Iteration. The biofilm evolution is calculated al-
ternating the computation of cellular processes up
to a time tC (estimated from the doubling time),
with steps of the water absorption processes in a
time scale tW ∼ tC (estimated from biomass pro-
duction) and steps of the deformation processes in a
time scale tD (which requires precise measurements
of the substratum parameters).
Summarizing, the mechanical processes change the ge-
ometry of the biofilm. The equations for the differ-
ent concentrations must be solved in the new geometry,
which affects the cellular processes altering the probabili-
ties for the different behaviors. The cellular processes al-
ter the elastic parameters and the pressure in the biofilm,
contributing residual stresses for its deformation due to
growth, death and fluid migration.
VI. SIMULATION RESULTS
Series of simulations were performed to illustrate the
behavior of the model and its limitations, choosing sin-
gle mechanisms or combinations of a few of them. The
results show wrinkled patterns and cell distributions in
qualitative agreement with recent experimental observa-
tions. They provide insight on the influence of parame-
ters, cellular activities and mechanical processes on the
biofilm shape and structure.
Basic growth and spread mechanisms do not produce
wrinkled shapes. Even if we activate water absorption
processes, the biofilm will spread faster, but no wrinkles
appear. When elastic deformation mechanisms are incor-
porated, wrinkles begin to form. We revisit the simula-
tion described in Figure 7 for the same parameter values
and a similar initial biofilm seed containing five peaks,
see Figure 11(a). During the third step of the growth,
spread and differentiation processes, EPS producers ap-
pear. At the fourth step, we consider that enough matrix
has been produced to regard the biofilm as an elastic film
with Young modulus E. During the 10-th step, dead cells
appear at the bottom of the initial peaks. Dead areas ex-
pand during the 11-th step. In the 12-th step, growth in
the central region has slowed down due to depleted nu-
trient levels and increased waste presence.
Figures 11(b)-(f) illustrate the spatial structure of the
corresponding growth tensor. Initial peaks diffuse as the
biofilm grows and become depressions when dead cells ap-
pear. Once growth in the core is depleted, compression is
higher in the border regions. A randomly perturbed ini-
tial biofilm deforms according to Eqs. (23)-(24) generat-
ing small wrinkles that coarsen as time evolves. Figures
11(g)-(k) represent the out-of-plane deformation of the
film at selected steps of the growth, spread and differen-
tiation processes. After each step, the average radius of
the biofilm increases about one cell. The deformation of
the biofilm is then calculated for a time τ
24 s. As wrin-
kles develop, the growth processes must be implemented
15
FIG. 11. (Color online) (a) Initial biofilm seed. (b), (c), (d),
(e), (f) Residual stresses ε0
xx averaged after 100 trials of the
stochastic processes at steps 4, 8, 10, 11, 12. The average
compression is −0.132, −0.1425, −0.146, −0.1437, −0.124.
yy has a similar structure and average. ε0
respectively. ε0
xy
takes smaller values and is neglected. (g), (h), (i), (j), (k)
Vertical displacements. The height mark dx corresponds to
one cell. Parameter values are γ = 8, ν = 0.5, E = 25 kP a,
νv = 0.45 and µv = 0.
in biofilms with a wavy bottom, contributing additional
spatial variations to the growth tensor (higher compres-
sion in the valleys, lower compression in the peaks).
Unless we activate the water absorption mechanism, as
in Figure 9, a necrotic region will develop in the biofilm
core. Water absorption facilitates nutrient diffusion and
waste removal, maintaining the cell normal functions.
In the computations, we have approximated the resid-
ual stresses ε0
yy by their constant average, mod-
ified by peaks or depressions that represent increased or
xx and ε0
decreased compression in certain areas. More accurate
smooth approximations can be automatically produced
using denoising strategies borrowed from image process-
ing [73]. As time evolves, singularities may develop due
to the presence of 'hanging' cells in the border. Such cells
are poorly connected to the rest of the biofilm. This ar-
tifact is solved by discarding such cells when computing
deformations. For the selected parameter values, biofilms
are roundish. Therefore, we may alternatively smooth
the biofilm border using the averaged support obtained
when averaging trials of the stochastic processes to iden-
tify spatial variations of the growth tensor. Wavy borders
like the ones observed in Figure 6 are usually found for
large values of kna2
. This number depends on the nutri-
DnKn
ent source and the bacterial strain, but tends to be small
( << 1) in practice.
What causes the successive wrinkle branching and the
radial branching observed in Figures 1(a) and 2(a)? As
we have noticed, EPS production gives the cellular ag-
gregate a certain cohesion. As shown by Figures 10 and
11, an expanding biofilm is under compression due to cell
division, EPS production and water absorption. The de-
velopment of wrinkles in Figure 11 is limited by size con-
siderations. The radius of the biofilm increases from 60
to 80 cells, it thickness is about 10 cells and γ = 8. Dou-
bling the biofilms maximum radius, and consequently γ,
it appears that a simple round compression front expand-
ing at a certain speed may produce wrinkle branching,
structured differently depending on the front speed and
the compression magnitude.
Motivated by previous observations,
the residual
stresses ε0(x, t) are assumed to behave like a radial front
ε0(r−vt) expanding at a certain speed v, r being the dis-
tance to the seed center. This allows to lower the compu-
tational cost. Figures 2(b) and 8 are generated deforming
circular biofilm seeds according to Eqs. (23)-(24) for such
residual stresses. The evolution starts from a configura-
tion with small random vertical displacements. Figure 8
takes ε0(r) = −0.1 constant in the central region. Then,
it decreases sharply. The front advances one tile every
14/τ s. Figure 12 shows the time evolution of the wrin-
kled area. Small wrinkles similar to those in Figure 11
form that coarsen as shown in those images. Figure 8 is a
three dimensional view of the two dimensional projection
depicted in Figure 12(d). The biofilm contains scattered
dead spots where the residual stresses decrease by a cer-
tain factor, affecting the way wrinkles nucleate. Once
the wrinkled area attains a certain extension, the height
of the wrinkles decreases unless we vary the compression
magnitude. If we wish to increase the wrinkle branching
rate, while maintaining or enhancing their height as the
wrinkled region expands, the magnitude of the compres-
sion has to increase radially. Figure 2(b) was computed
raising the speed to one tile every 1.4/τ s and increasing
the compression magnitude by r/80 as the radius of the
compressed region grows. When the compression front
expands too slowly, rings may form around the wrinkled
area. If it expands too fast, we may see different types
16
τ , 560 T
τ , 1120 T
FIG. 12. (Color online) Snapshots of the formation of wrin-
kles for ε0 = 0.1.
(a), (b), (c), (d) images taken at times
280 T
τ , 840 T
τ . An initial random state coarsens
to produce a connected wrinkle network that opens up into
radial branches as the compressed region spreads. The biofilm
has Poisson ratio ν = 0.5 and Young modulus E = 25 kP a.
The Poisson ratio and rubbery modulus of the substratum are
νv = 0.45 and µv = 0. T and τ are defined in Eq. (25). The
ratio of the in-plane spatial scale to the out-of-plane spatial
scale is set to γ = 16. The equations are nondimensionalized
so that the dimensionless biofilm thickness becomes h = 1.
τ dx2. The spatial step is dx = 0.1h.
The time step is dt = 0.14
of geometric shapes. In between, the network of wrinkles
opens up radially as it expands. The outer branches are
connected to the inner network and interact with it.
So far, we have discussed branching of wrinkles. What
produces the wrinkled corona in Figure 1(a)? A sud-
den stop of an advancing compression front may arrange
branching wrinkles in a corona type structure temporar-
ily. The whole network changes its structure later. A uni-
formly compressed film whose elastic modulus decreases
in an outer corona develops stable radial wrinkles, as in
Figure 1(b). The Young modulus E has been decreased
by a factor 0.5 in the outer corona. Radial residual com-
pression of constant magnitude −0.1 is applied every-
where, for the same parameter values as in Figures 8
and 12. Now, there is no advancing front. The biofilm
boundary remains fixed and the residual stresses are con-
stant everywhere. Radial decrease of the Young modulus
might be due to larger water absorption rates in the outer
regions, presumably softer, due to the larger presence of
normal cells.
Too sharp a decrease in the elastic modulus, for the se-
lected parameters, combined with smaller values of Pois-
son's ratio, may result in fast coarsening and disappear-
ance of the central wrinkles, as in Figure 13(a). The
compression factor is kept constant in a central core but
increases radially in an outer corona. The inner network
of wrinkles vanishes with time, but a corona of increas-
ing radial wrinkles is formed. Both the Young modulus
E and the Poisson ratio ν affect the time scales. Wrinkles
coarsen faster in regions with higher Young modulus or
lower Poisson ratio. Central wrinkles, however, are an-
chored by the presence of a couple of dead spots in Fig-
ure 13(b). Decreasing the Poisson ratio enhances height.
Lowering the Poisson's ratio to 0.3, the outer wrinkles
not only maintain their heights, but also may increase it
for increasing radial compression. Unlike Figure 1(b) or
Figure 8, a radially increasing compression factor ε0 pro-
duces an inner network of wrinkles that opens up form-
ing branches whose height can be kept uniform over the
biofilm, see Figure 13(c). Notice also that γ is decreased
by a half. Radial increase of the compression might be
due to larger growth rates in the outer regions caused by
increasing availability of nutrients.
VII. CONCLUSIONS
Replicating cell populations create three dimensional
organisms of diverse shapes. Elucidating how those dif-
ferent geometries arise is an intriguing question that has
motivated many theories. Small cellular systems, such as
bacterial biofilms, seem to develop from the interplay be-
tween cellular and mechanical processes. Studying the in-
teraction of those two mechanisms from an experimental
point of view requires concurrent measurement of both
processes, posing a major challenge. Developing mathe-
matical and computational frameworks able to incorpo-
rate the increasing amount of experimental observations
in their governing rules may provide insight on the feed-
back between microscopic cell behavior and macroscopic
continuum processes. We have proposed a hybrid model
to describe the growth dynamics of a small cluster of
biofilm on an air-agar interface. We are able to trans-
fer information from a stochastic description of cellular
processes into a continuum model for the deformation
of the film, and use these deformations as a feedback
for the cell activities. We have shown that mechanical
processes change the geometry of the biofilm. The rele-
vant concentrations must be computed in a new geome-
try. This fact influences the cellular processes, modifying
the probabilities for the cell behaviors. In turn, the cel-
lular processes modulate the elastic parameters and the
pressure in the biofilm, and generate residual stresses due
to death, growth, and fluid migration. This microscopic
feedback determines the subsequent mechanical processes
and so on. We have observed that the properties of the
substratum together with the mechanical properties of
the biofilm and residual stresses due to death, growth
and swelling seem to control wrinkled biofilm shapes.
Our in silico biofilms agree qualitatively with some ex-
perimental observations, in the sense that their shape
and the spatial distribution of the different types of dif-
ferentiated cells is similar. Biomass, air, water and agar
distribute through the tiles of a computational grid. We
confer tiles occupied by alive cells the ability to change
17
FIG. 13. (Color online) Wrinkled structures obtained vary-
ing the residual stresses and the elastic moduli for a smaller
film with lower Poisson's ratio. (a) Transition from a harder
central region with larger Young modulus to a softer outer
corona: radial branches in the corona are separated from the
core by rings. (b) Uniform residual compression in the cen-
tral region with localized sinks due to dead spots plus radially
increasing compression in the outer corona: small wrinkles an-
chored by the dead spots in the center merge with larger ra-
dial wrinkles in the outer corona. (c) Residual compression in-
creasing slowly with distance to the center: labyrinths formed
in the center become radial branches in the outer corona. Pa-
rameter values: Same as previous figures except ν = 0.3 and
γ = 8. dx is the spatial step.
their status according to probability laws informed by a
cascade of concentration fields: nutrients, waste, ComX
and surfactin. Local inhomogeneities trigger cell death at
the bottom of biofilm peaks. Nutrient depletion deacti-
vates cells at the top. When a threshold level of ComX is
reached, surfactin producers appear. When a threshold
level of surfactin is achieved, EPS producers proliferate.
Once EPS release activates, we consider the biofilm an
elastic film, whose elastic parameters are modulated by
its microstructure. We also launch a mechanism for wa-
ter absorption triggered by variations in osmotic pressure
caused by EPS production. Weighting at each biofilm lo-
cation the average values of elastic moduli taken from
experiments according to the status of the neighboring
tiles, we may produce spatially varying parameters tak-
ing into account the microstructure. The biofilm weak-
ens due to the presence dead cells.
It also softens in
swollen regions, due to water absorption. Biofilm expan-
sion is quantified by means of a growth tensor computed
from the stochastic growth, death and water absorption
processes. In this way, we estimate the residual stresses
caused by these mechanisms. The compression residual
stresses and the spatially varying moduli used in the nu-
merical tests presented here are motivated by those com-
putations. We study biofilm deformation in response to
those stresses by means of a Foppl-Von K´arm´an approx-
imation, obtaining intricate wrinkled cores that split in
radial branches. Water erodes the regions were cells die,
expanding along the wrinkles. This improves transport
of waste and nutrients, hindering cell death and favoring
growth.
18
outer corona, also enhance the outer radial wrinkles.
Quantitative comparison with experiments should re-
quire careful calibration of several parameters, in par-
ticular, regarding substratum properties and heuristic
probability laws. Substratum parameters such as its
Poisson ratio, thickness, viscosity and rubbery modulus
should be accurately measured since they affect the time
scales for the biofilm dynamics. Uptake rates, produc-
tion rates, saturation constants and diffusivities of the
involved chemical compounds should be determined too,
since they influence the extend of growth, death and dif-
ferentiation processes. As the biofilm thickens and its
spatial heterogeneity increases, the Foppl-Von K´arm´an
approximation should likely be replaced by a fully three
dimensional elasticity model. The spread mechanism and
the dynamics of water in the system should be revised
too.
ACKNOWLEDGMENTS
Our simulations show that wrinkling seems to be as-
sociated with stiffness fluctuations. Variations in inner
residual stresses and elastic constants due to growth,
swelling and death seem to govern wrinkle nucleation and
branching. Compression stresses expanding radially due
to swelling and growth produce wrinkled cores that split
in radial wrinkles. The presence of dead regions alters
the way the core wrinkles are nucleated. It also favors
wrinkle formation and persistence around dead areas as
the compression rate is reduced or time increases. Suc-
cessive wrinkle branching may occur depending on the
expansion velocity. Radially graded compression stresses
may enhance the outer radial wrinkles. Spatially depen-
dent elastic moduli, harder at the center and softer in the
A. Carpio and D.R. Espeso thank M.P. Brenner for
hospitality while visiting Harvard University. D. R. Es-
peso and A. Carpio were supported by Comunidad de
Madrid and the spanish MICINN through grants No.
S2009/DPI-1572, No. FIS2011-28838-C02-02 and No.
FIS2010-22438-E. A. Carpio was also supported by a mo-
bility grant of Fundaci´on Caja Madrid. B. Einarsson was
supported by the NILS Mobility project (European eco-
nomic area-EEA grant) and MICINN grant No. FIS2008-
04921-C02-01. Part of the computations of this work
were performed in EOLO, the HPC of Climate Change
of the International Campus of Excellence of Moncloa,
funded by MECD and MICINN. This is a contribution
to CEI Moncloa.
[1] R.D. Monds and G.A. O'Toole, Trends Microbiol. 17, 73
[12] B. Purevdorj, J.W. Costerton and P. Stoodley, Appl. En-
(2009).
viron. Microbiol. 68, 4457 (2002).
[2] H.C. Flemming and J. Wingender, Nature Rev. Micro-
[13] K. Drescher, Y. Shen, B.L. Bassler and H.A. Stone, Proc.
biol. 8, 623 (2010).
Nat. Acad. Sc. 110, 4345 (2013).
[3] N. Hoiby, T. Bjarnsholt, M. Givskov, S. Molin and O.
[14] L. Chai, H. Vlamakis and R. Kolter, MRS Bulletin 36,
Ciofu, Int. J. Antimicrob. Agents 35, 322 (2010).
374 (2011).
[4] J.W. Costerton, P.S. Stewart and E.P. Greenberg, Sci-
[15] J.N. Wilking, T.E. Angelini, A. Seminara, M.P. Brenner
ence 284, 1318 (1999).
[5] R.O. Darouiche, Clin. Infect. Dis. 33, 1567 (2001).
[6] C.C. de Carvalho, Recent Pat. Biotechnol. 1, 49 (2007).
[7] E. Eguia, A. Trueba, B. Ro-Calonge, A. Giron, C. Bielva,
Int. Biodeter. Biodeg. 62, 79 (2008).
[8] Y. Xiong, Y. Liu, Appl. Microbiol. Biotechnol. 86, 825
(2010).
[9] R. Singh, D. Paul and R.K. Jain, Trends Microbiol. 14,
389 (2006).
[10] B. Schachter, Nature Biotechnol. 21, 361 (2003).
[11] S. Wuertz, P.L. Bishop and P.A. Wilderer, Biofilms
in wastewater treatment: An interdisciplinary approach
(IWA Publishing, 2003).
and D.A. Weitz, MRS Bulletin 36, 38 (2011).
[16] M. Trejo, C. Douarche, V. Bailleux, C. Poulard, S. Mar-
iot, C. Regeard and E. Raspaud, Proc. Nat. Acad. Sc.
110, 2011 (2013).
[17] M. Asally, M. Kittisopikul, P. Ru´e, Y. Du, Z. Hu, T.
C¸ agatay, A.B. Robinson, H. Lu, J. Garcia-Ojalvo and
G.M. Suel, Proc. Nat. Acad. Sc. 109, 18891 (2012).
[18] E. Coen, A.G. Rolland-Lagan, M. Matthews, J.A. Bang-
ham and P. Prusinkiewicz, Proc. Nat. Acad. Sci. 101,
4728 (2004).
[19] T. Mammoto and D.E. Ingber, Development 137, 1407
(2010).
[20] O. Wanner, H. Eberl, E. Morgenroth, D. Noguera, C.
Picioreanu, B. Rittmann, and M.C.M. van Loosdrecht,
19
Mathematical modeling of biofilms (IWA Publishing, Lon-
don UK, 2006).
[21] I. Klapper and J. Dockery, SIAM Rev. 52, 221 (2011).
[22] J.B. Xavier and K.R. Foster, Proc. Nat. Acad. Sc. 104,
876 (2007).
[23] J.B. Xavier, E. Martinez-Garcia, and K.R. Foster, The
American Naturalist 174, 1 (2009).
cations, Computation, and Finance, 2003, The IMA Vol-
umes in Mathematics and its Applications, Vol. 134, p.1.
[49] C.S. Laspidou, L.A. Spyrou, N. Aravas and B.E.
Rittmann, Math. Biosci. 251, 11 (2014).
[50] M.H. Swat, G.L. Thomas, J.M. Belmonte, A. Shirinifard,
D. Hmeljak and J.A. Glazier, Methods Cell Biol. 110,
325 (2012).
[24] D. Rodriguez, B. Einarsson and A. Carpio, Phys. Rev. E
[51] T. Storck, C. Picioreanu, B. Virdis and D.J. Batstone,
86, 061914 (2012).
Biophys. J. 106, 2037 (2014).
[25] E. Alpkvist, C. Picioreanu, M.C.M. Loosdrecht and A.
[52] P. Mehta, R. Mukhopadhyay and N.S. Wingreen, Phys.
Heyden, Biotechnol. Bioeng. 94, 961 (2006).
Biol. 5, 026005 (2008).
[26] T. Shaw, M. Winston, C.J. Rupp, I. Klapper and P.
[53] W.C. Fuqua, S.C. Winans and E.P. Greenberg, J. Bac-
Stoodley, Phys. Rev. Lett. 93, 098102 (2004).
teriol. 176, 269 (1994).
[27] N.G. Cogan and R.D. Guy, HFSP J. 4, 11 (2010).
[28] H.J. Dupin and P.K. Kitanidis, P.L. McCarty, Water Res.
37, 2965 (2001).
[29] D. Taherzadeh, C. Picioreanu, U. Kuttler, A. Simone,
W.A. Wall and H. Horn, Biotech. Bioeng. 105, 600
(2009).
[30] N. Autrusson, L. Guglielmini, S. Lecuyer, R. Rusconi,
and H.A. Stone, Phys. Fluids 23, 063602 (2011).
[31] S.W. Hermanowicz, Math. Biosci. 169, 1 (2001).
[32] C.S. Laspidou and B.E. Rittmann, Water Res. 38, 3349
[54] B.A. Hense, C. Kuttler, J. Muller, M. Rothballer, A.
Hartmann and J.U. Kreft, Nature Rev. Microbiol. 5, 231
(2007).
[55] T. Long, K.C. Tu, Y. Wang, P. Mehta, N.P. Ong, B.L.
Bassler and N.S. Wingreen, PLOS Biol. 7, e1000068
(2009).
[56] S.W. Teng, J.N. Schaffer, K.C. Tu, P. Mehta, W. Lu,
N.P. Ong, B.L. Bassler and N.S. Wingreen, Mol. Syst.
Biol. 7, 491 (2011).
[57] P.S. Stewart and M.J. Franklin, Nature Rev. Microbiol.
(2004).
[33] G.E. Kapellos, T.S. Alexiou and A.C. Payatakes, Adv.
Water Resour. 30, 1648 (2007).
[34] N.J. Poplawski, A. Shirinifard, S. Macie and J.A. Glazier,
Math. Biosc. Eng. 5, 355 (2008).
[35] L.A. Lardon, B.V. Merkey, S. Martins, A. Dotsch, C. Pi-
cioreanu, J.U. Kreft and B.F. Smets, Environ. Microbiol.
13, 2416 (2011).
[36] A. Seminara, T.E. Angelini, J.N. Wilking, H. Vlamakis,
S. Ebrahim, R. Kolter, D.A. Weitz and M.P. Brenner,
Proc. Nat. Acad. Sc. 109, 1116 (2012).
[37] J.N. Wilking, V. Zaburdaev, M. De Volder, R. Losick,
M.P. Brenner and D.A. Weitz, Proc. Nac. Acad. Sc. 110,
848 (2013).
[38] M. Ben Amar and M. Wu, EPL 108, 38003 (2014).
[39] T. Mora and A. Boudaoud, Eur. Phys. J. E 20, 119
6, 199 (2008).
[58] T.Y. Hou, Acta Numerica 4, 335 (1995).
[59] W. Zhang, A. Seminara, M. Suaris, M.P. Brenner, D.A.
Weitz and T.E. Angelini, New J. Phys. 16, 015028
(2014).
[60] P. Mehta, S. Goyal, T. Long, B.L. Bassler and N.S.
Wingreen, Mol. Syst. Biol. 5, 325 (2009).
[61] L.R. Swem, D.L. Swem, N.S. Wingreen and B.L. Bassler,
Cell 134, 461 (2008).
[62] C.D. Nadell, J.B. Xavier, S.A. Levin and K.R. Foster,
PLOS Biol. 6, e14 (2008).
[63] M. Hendrata and B. Birnir, Phys. Rev. E 81, 061902
(2010).
[64] Z. Hejnowicz and J.A. Romberger, J. Theor. Biol. 110,
93 (1984).
[65] T. Glozman, H. Azhari, J. Ultrasound Med. 29, 387
(2006).
(2010).
[40] R. Huang and S.H. Im, Phys. Rev. E 74, 026214 (2006).
[41] Z. Liu, S. Swaddiwudhipongc and W. Hongd, Soft Matter
9, 577 (2013).
[66] K.S. Anseth, C.N. Bowman and L Brannon-Peppas, Bio-
materials 17, 1647 (1996).
[67] K.A. Ross and M.G. Scanlon, J. Texture Stud. 30, 17
[42] Y. Ni, D. Yang and L. He, Phys. Rev. E 86, 031604
(1999).
(2012).
[68] S. Istini, M. Ohno and H. Kusunose, Bull. Mar. Sci. Fish.
[43] L.D. Landau and E.M. Lifschitz, Theory of elasticity
Kochi Univ. 14, 49 (1994).
(Pergamon Press, 1986).
[44] J. Dervaux and M. Ben Amar, Phys. Rev. Lett. 101,
068101 (2008).
[69] E. Fernandez, D. Lopez, C. Mijangos, M. Duskova-
Smrckova, M. Ilavsky and K. Duseck, J. Polym Sci: Part
B: Polym. Phys. 46, 322 (2008).
[45] J. Dervaux, P. Ciarletta and M. Ben Amar, J. Mech.
[70] M. Ahearne, Y. Yang, A.J. El Haj, K. Then and K.K.
Phys. Solids 57, 458 (2009).
Liu, J. R. Soc. Interface 2, 455 (2005).
[46] M.A.A. Grant, B. Waclaw, R.J. Allen and P. Cicuta, J.R.
[71] G.E. Kapellos, T.S. Alexiou and A.C. Payatakes, Math.
Soc. Interface 11, 20140400 (2014).
Biosci. 225, 83 (2010).
[47] A. Dechesne, G. Wang, G. Gulez, D. Or and B.F. Smets,
[72] G.E. Kapellos, T.S. Alexiou and A.C. Payatakes, Int. J.
Proc. Nat. Acad. Sc. 107, 14369 (2010).
Eng. Sc. 51, 241 (2012).
[48] M.S. Alber, M.A. Kiskowski, J.A. Glazier and Y. Jiang,
in Mathematical Systems Theory in Biology, Communi-
[73] A. Torres, A. Marquina, J.A. Font, J.M. Ib´anez, Phys.
Rev. D 90, 084029 (2014).
|
1204.4361 | 2 | 1204 | 2012-05-17T12:33:55 | Super-sensitive Molecule-hugging Graphene Nanopores | [
"physics.bio-ph",
"cond-mat.soft",
"physics.ins-det"
] | Longitudinal resolution and lateral sensitivity are decisive characteristics that determine the suitability of a nanopore sensor for sequencing a strand of DNA as well as other important polymers. Previous modeling of DNA induced ionic current blockades in single atom thick graphene nanopores has shown these nanopores to have sufficient longitudinal resolution to distinguish individual nucleobases along the length of a DNA molecule. Here we experimentally focus on the sensitivity to small changes in DNA diameter that can be discerned with graphene nanopores. We show that remarkably large sensitivities (0.5 nA/A)are obtained when the nanopore is tailored to have a diameter close to that of the polymer of interest. Our results have been obtained with double-stranded DNA (dsDNA). Smaller graphene nanopores that can be tuned to the diameter of single-stranded DNA (ssDNA) for sequencing have only recently been demonstrated. Our results indicate that nanopore sensors based on such pores will provide excellent resolution and base discrimination, and will be robust competitors to protein nanopores that have recently demonstrated sequencing capability. | physics.bio-ph | physics | Super-sensitive Molecule-hugging Graphene Nanopores
Slaven Garaj1, Song Liu1,2, Daniel Branton3, and Jene A. Golovchenko1,4*
1Department of Physics, Harvard University, Cambridge Massachusetts, 02138, USA
2Department of Physics, Peking University, Beijing 100871, P.R. China
3Department of Molecular and Cellular Biology, Harvard University, Cambridge, Massachusetts, 02138, USA
4School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts, 02138, USA
*Corresponding author, email: [email protected]
1
Longitudinal resolution and lateral sensitivity are decisive characteristics that determine
the suitability of a nanopore sensor for sequencing a strand of DNA as well as other
important polymers. Previous modeling of DNA induced ionic current blockades in single
atom thick graphene nanopores has shown these nanopores to have sufficient longitudinal
resolution to distinguish individual nucleobases along the length of a DNA molecule1. Here
we experimentally focus on the sensitivity to small changes in DNA diameter that can be
discerned with graphene nanopores. We show that remarkably large sensitivities
(0.5 nA/Å) are obtained when the nanopore is tailored to have a diameter close to that of
the polymer of interest. Our results have been obtained with double-stranded DNA
(dsDNA). Smaller graphene nanopores that can be tuned to the diameter of single-
stranded DNA (ssDNA) for sequencing have only recently been demonstrated2. Our
results indicate that nanopore sensors based on such pores will provide excellent resolution
and base discrimination, and will be robust competitors to protein nanopores that have
recently demonstrated sequencing capability3,4.
Biological5 and solid-state6 nanopore sensors are becoming key elements in next
generation DNA sequencing technologies7,8. They are increasingly being used for rapid
characterization of many individual DNA molecules9-11, proteins12,13, and protein-DNA
complexes14,15. In a nanopore sensor, a charged biopolymer in solution is electrophoretically
driven through a nanometer scale pore in an insulating membrane separating two voltage biased
reservoirs filled with aqueous ionic solution. As the negatively charged DNA molecules pass
through a nanopore, they block the flow of ions through the nanopore, leading to a transient drop
of the ionic current through the nanopore (a blockade). The instantaneous ionic current during
2
the blockade is related to the geometrical and chemical properties of the pore and the part of the
polymer in the pore at that given time. Here we demonstrate an extraordinary sensitivity of
graphene nanopores to DNA molecules when the pore and molecule diameters are very closely
matched.
We fabricated single nanopores in graphene membranes1, see also16,17, with a focused
beam of a 200 KeV transmission electron microscope (TEM). The graphene membranes were
suspended over 200 nm x 200 nm apertures in thin, free-standing SiNx films supported on silicon
frames (Fig. 1a and Methods section). DNA translocation through these graphene nanopores
was investigated in high salt and high pH solutions (3M KCl, pH 10). Under these conditions
electrostatic and chemical stick-slip forces between DNA molecules and the graphene surface are
largely suppressed as will be demonstrated below. The applied voltage bias between the
reservoirs on either side of the graphene membrane was 160 mV.
Figure 1b shows individual current blockades in graphene nanopores of various diameters
induced by 10kb dsDNA molecules. A strong nanopore diameter dependence is evident, with
c) or protein nanopores. This is especially clear for the nanopores whose diameters approach the
blockade currents that are much larger than have been observed in SiNx nanopores (Fig. 1b and
size of the translocating polymer. As we show below, such polymer-hugging nanopores are
extraordinarily sensitive to small changes in the difference between the molecule and nanopore
diameter.
Figure 1d shows ionic conductivity vs. nanopore diameter for a series of open graphene
nanopores. The linear dependence of conductivity on diameter is expected only when the
membrane is much thinner than the pore diameter1. This is consistent with the predicted access
3
resistance of a small circular pore18, our numerical solutions of the Laplace equation for the ionic
current density1, and molecular dynamic calculations19.
To understand the sensitivity of graphene nanopores to single molecule DNA
translocation, we investigated many DNA translocation events for pores of different diameters
(Figs. 2a & b). In nanopores with diameters greater than ~4 nm, dsDNA molecules translocate
through the pore either as extended linear molecules (leading to a single-dip event, Fig. 2c, left
the blockade (Fig. 2c, right trace). In nanopores with diameters ≤3.2 nm, only unfolded
trace) or as a folded molecule (typically doubling the current blockade during the folded part of
molecules were observed (single-dip events), consistent with the notion that these nanopores are
too small to admit folded dsDNA molecules for passage. We characterize each DNA
translocation event by two parameters: the event duration, which indicates the time it takes for
the molecule to fully translocate through the nanopore, and the average current blockade
magnitude, which reflects the extent to which the molecule obstructs or blocks the flow of ionic
current through the nanopore. The product of these two parameters is the electronic charge
deficit10 (ecd) during the event. Note the clustering of all the dsDNA translocation events
around the hyperbolic line of constant ecd in Fig. 2a. This, together with the short average
translocation time (Tm~150μs), shows that dsDNA molecules in 3M KCl, pH 10, freely
translocate through the graphene nanopore and are minimally retarded by stochastic sticking
interactions with the graphene membrane or nanopore10.
Raising the pH of the salt solution from 10 to 12.5 denatures the dsDNA into single-
stranded DNA (ssDNA) molecules11. Unfolded ssDNA translocation events (Fig. 2d right) show
blockade magnitudes less than half that of dsDNA (Fig. 2d, left trace), a reflection of the reduced
diameter of ssDNA. To our knowledge this is the first observation of ssDNA translocating
4
through a graphene nanopore. Many ssDNA events are observed far from the line of constant
ecd because of their extended translocation times (Fig. 2c). This strongly suggests that many of
the ssDNA molecules stick to, or interact, with graphene. This is not surprising since the
nucleobases in ssDNA are accessible and known to form π-π stacking/van der Waals interactions
with a graphene surface20.
The average current blockade for unfolded dsDNA translocations in graphene nanopores
was measured as a function of nanopore diameter, and the results are shown as the filled circles
in Fig. 3a. A steep increase in blockade current with decreasing nanopore diameter, from 6.5 to
2.7 nm, is observed. This indicates a continuing increase in sensitivity to the difference in
molecule diameter and pore diameter as the fit between the two gets closer and closer. The open
circles are data obtained from e-beam drilled nanopores in 30 nm thick SiNx which, because of
the membrane thickness, show virtually no change in blockade current with nanopore diameter.
To understand the experimental data in Fig. 3a, we calculated (see Methods) current flow
through an open pore (I0) vs. pore diameter, and through the same pore containing a DNA
molecule passing through its center, to obtain the blockade current (IB). The electrolyte solution
is modeled as a continuous medium with appropriate ionic conductivity, and we numerically
solved the Laplace equation with appropriate boundary conditions. We fitted the results to the
experimental data using the dimensions of the boundary conditions as the fitting parameters. We
call DIC the ionic conducting diameter of the pore,
ITL the membrane insulating thickness and
DNAd
the diameter of the insulating DNA molecule at the center of the pore. Actually, rather
than fit the value of each pore diameter we set the effective diameter of each nanopore DIC =
Dpore – δDIC, to be reduced from each TEM measured pore diameter
poreD by δDIC . The latter is
a single fitting parameter used for all pores. This correction allows for the fact that solution ions
5
do not approach the hydrophobic TEM determined pore perimeter due to various chemical size
effects. The best fit to the data in Fig. 3a is given by the solid line for which δDIC =( 0.65 ±
0.05) nm ,
DNAd
1.93 0.02 nm
, and LIT =( 0.37 ± 0.08) nm. The value of LIT determined here
is in the range of our previously reported value (0.6 [+0.9, –0.6] nm)1 and agrees with
theoretical predictions21 for a graphene-water distance of 0.31-0.34 nm. It seems reasonable that
the effective diameter reduction δDIC be similar to the membrane insulating thickness LIT with
any deviations attributed to the energy cost of an ion shedding its hydration shell within the
nanopore22,23, or the consequences of chemical groups bound to the graphene nanopore’s edge22.
The modeled dependence of blockade current vs. pore diameter is shown as the solid
curve in Fig. 3a based on the parameters given above. The fit to the data (solid circles) is seen
to be rather good and we conclude that the calculation captures the essential physics with the
parameters LIT and DIC absorbing all the molecular interaction effects24.
We define the nanopore’s sensitivity as the change in current blockade caused by a
S ( DIC , d poly )
change in the diameter of the translocating polymer
This is the slope of the solid curve as a function of nanopore diameter. The approximation
assumes dpoly DIC > LIT.
I B
d poly
I B
DIC
The value of S from Fig. 3a, remarkably, exceeds 0.5 nA/Å, predicting that a change of a
tenth Angstrom molecule diameter in our close fitting pore results in a current blockade change
of 50 pA in BI for dsDNA. Similar results should be obtained for ssDNA (
d
1.4
nm
) with a
similarly tight fitting nanopore.
6
Figure 3b shows the current density in and near a 2.5 nm DNA threaded pore using the
relevant fitting parameters. The high sensitivity is due to the very concentrated current density in
the narrow solution space between the DNA and nanopore periphery. Figure 3c shows a plot of
the Sensitivity Gradient along the DNA molecule surface defined by
where D is the polymer diameter, R the pore resistance, z the distance from the symmetry plane
1
lim
D R
D
0
dV
D
dz
dV
D D
dz
dS
dz
of the pore along the surface of the polymer, and
DV z is the electrostatic potential along the
( )
surface of a polymer of diameter D as a function of distance from the pore z. The Sensitivity
Gradient identifies where the contributions to S come from along the molecule’s length and thus
reveals the spatial resolution of the molecule pore system. The Sensitivity Gradient “full width
half max” of 0.5 nm in Figure 3c, for the 2.5 nm pore, is the order of expected base separations
in ssDNA. The excellent spatial resolution along the molecule is closely connected with the
rapid drop in current density on leaving the vicinity of the pore as seen in Fig. 3b.
The sensitivity to molecule size demonstrated here and the high spatial resolution along
the molecule’s length shown previously1 are the direct consequence of the atomic thinness of a
molecule-hugging graphene nanopore. This, together with the recent proof-of-principle
demonstration that very small diameter graphene nanopores can be fabricated with atomic-scale
precision2, suggest that graphene nanopores will be an ideal nanopore sensor.
Achieving ssDNA strand sequencing with graphene nanopores will also require solutions
to several of the same problems that have been addressed when utilizing protein pores: the need
for ratcheting control of DNA translocation to suppress Brownian molecular motions25, and the
suppression of large 1/f electronic noise that can severely degrade the ionic current signal-to-
7
noise. It remains to be seen if some of the solutions for translocation control used with
biological pores3 may be applied to graphene pores or otherwise engineered26. Likewise, further
research will be required to determine how to optimally suppress the large observed 1/f
electronic noise frequently observed in graphene (Fig. 1e). This noise can be reduced by
applying various coatings16,17, but these coatings increase the length of the nanopore and are
incompatible with the high resolution detector strategy reported here. We have achieved noise
reduction by simply reducing the free standing area of the graphene membrane (Fig. 1e).
In conclusion we believe the longitudinal sub-nanometer resolution possible in graphene
nanopores1 together with the radial super-sensitivity discovered here bode well for the
application of graphene nanopores to many molecular sensing, characterization, and sequencing
problems.
Methods
The graphene was grown via chemical vapor deposition (CVD) on a copper foil27, and
subsequently transferred28 over the aperture on a free-standing SiNx film covering a silicon
chip, and carefully processed to remove any residual hydrocarbon impurities from the polymer
support film used during the transfer process1. Individual graphene nanopores were drilled with
a JEOL 2010F TEM operating at 200kV. Micro-Raman measurements (model WITec CRM 200)
and TEM imaging with atomic resolution (aberration-corrected Zeiss Libra 200, operating at
80keV electron energies) demonstrated that the transferred graphene film was single-layer and
largely free of hydrocarbon contaminants. The graphene chip was sealed in a fluidic cell so as to
separate two chambers that were subsequently filled with 3 M potassium-chloride solutions, pH
10 or pH12.5. The graphene nanopore was the only path through which ions and DNA
8
molecules (10kBase fragments) could pass between the two chambers. Applying a constant
160 mV bias voltage with Ag/AgCl electrodes in each chamber, we measured the ionic current
and DNA translocation through the nanopore using standard electrophysiological methods.
The numerical simulations were performed using the COMSOL Multiphysics finite
element solver in 3D geometry, cylindrically symmetric along the axis of the nanopore. A DNA
molecule was modeled as a long stiff insulating rod threading the nanopore along its axis. We
solved the full set of Poisson–Nerst–Planck (PNP) equations, with the boundary conditions at the
graphene corresponding to idealized, uncharged membrane impermeable to ions. In our
experimental regime with high KCl concentration and small applied voltage, the PNP solution
was found to differ only by few percents from the numerical solution of the Laplace equation
with the appropriate electrolyte conductivity. The total ionic current was calculated by
integrating current density across the diameter of the nanopore.
References
1.
Garaj, S. et al. Graphene as a subnanometre trans-electrode membrane. Nature 467, 190-
194 (2010).
2.
Russo, C.J. & Golovchenko, J.A. Atom-by-atom nucleation and growth of graphene
nanopores. Proc. Natl. Acad. Sci. U.S.A. 109, 5953-5957 (2012).
3.
Cherf, G.M. et al. Automated forward and reverse ratcheting of DNA in a nanopore at 5-
Å precision. Nature Biotechnology 30, 344-348 (2012).
4.
Manrao, E.A. et al. Reading DNA at single-nucleotide resolution with a mutant MspA
nanopore and phi29 DNA polymerase. Nature Biotech. 30, 349-353 (2012).
9
5.
Kasianowicz, J.J., Brandin, E., Branton, D. & Deamer, D.W. Characterization of
individual polynucleotide molecules using a membrane channel. Proc. Natl. Acad. Sci.
U.S.A. 93, 13770-13773 (1996).
6.
7.
Li, J. et al. Ion-beam sculpting at nanometre length scales. Nature 412, 166-169 (2001).
Branton, D. et al. The potential and challenges of nanopore sequencing. Nature
Biotechnology 26, 1146-1153 (2008).
8.
Venkatesan, B.M. & Bashir, R. Nanopore sensors for nucleic acid analysis. Nature
Nanotechnology 6, 615-624 (2011).
9.
Deamer, D. & Branton, D. Characterization of nucleic acids by nanopore analysis. Acc.
Chem. Res. 35, 817-825 (2002).
10.
Li, J., Gershow, M., Stein, D., Brandin, E. & Golovchenko, J. DNA molecules and
configurations in a solid-state nanopore microscope. Nature Materials 2, 611-615 (2003).
11.
Fologea, D. et al. Detecting single stranded DNA with a solid state nanopore. Nano Lett.
5, 1905-1909 (2005).
12.
Han, A. et al. Sensing protein molecules using nanofabricated pores. Appl. Phys. Lett. 88,
093901 (2006).
13.
Talaga, D.S. & Li, J. Single-molecule protein unfolding in solid state nanopores. J. Am.
Chem. Soc. 131, 9287-9297 (2009).
14.
Olasagasti, F. et al. Replication of individual DNA molecules under electronic control
using a protein nanopore. Nature Nanotechnology 5, 798-806 (2010).
15.
Venkatesan, B.M. et al. Stacked graphene-Al2O3 nanopore sensors for sensitive detection
of DNA and DNA-protein complexes. ACS Nano 6, 441-450 (2012).
10
16. Merchant, C.A. et al. DNA translocation through graphene nanopores. Nano Lett. 10,
2915-2921 (2010).
17.
Schneider, G.F. et al. DNA translocation through graphene nanopores. Nano Lett. 10,
3163–3167 (2010).
18.
Hall, J.E. Access resistance of a small circular pore. The Journal of General Physiology
66, 531-532 (1975).
19.
Sathe, C., Zou, X., Leburton, J.-P. & Schulten, K. Computational investigation of DNA
detection using graphene nanopores. ACS Nano 5, 8842-8851 (2011).
20.
Gowtham, S., Scheicher, R.H., Ahuja, R., Pandey, R. & Karna, S.P. Physisorption of
nucleobases on graphene: Density-functional calculations. Phys. Rev. B 76, 033401
(2007).
21.
Alexiadis, A. & Kassinos, S. Molecular simulation of water in carbon nanotubes. Chem.
Rev. 108, 5014-5034 (2008).
22.
Sint, K., Wang, B. & Kral, P. Selective ion passage through functionalized graphene
nanopores. J. Am. Chem. Soc. 130, 16448-16449 (2008).
23.
Zwolak, M., Lagerqvist, J. & Di Ventra, M. Quantized ionic conductance in nanopores.
Phys. Rev. Lett. 103, 128102 (2009).
24.
Eijkel, J.C.T. & Berg, A.V.D. Nanofluidics: what is it and what can we expect from it?
Microfluidics and Nanofluidics 1, 249-267 (2005).
25.
Lu, B., Albertorio, F., Hoogerheide, D.P. & Golovchenko, J.A. Origins and consequences
of velocity fluctuations during DNA passage through a nanopore. Biophys. J. 101, 70-79
(2011).
11
26.
Luan, B. et al. Base-by-base ratcheting of single stranded DNA through a solid-state
nanopore. Phys. Rev. Lett. 104, 238103-238101 - 238103-238104 (2010).
27.
Li, X. et al. Large-area synthesis of high-quality and uniform graphene films on copper
foils. Science 324, 1312-1314 (2009).
28.
Lin, Y.-C. et al. Clean transfer of graphene for isolation and suspension. ACS Nano 5,
2362-2368 (2011).
12
Acknowledgements
This work was funded by a grant (number R01HG003703) to J.A. Golovchenko and D.
Branton from the National Human Genome Research Institute, National Institutes of Health. S.
Liu acknowledges support from the State Scholarship Fund of China.
Author Contributions
The experiments were designed by S.G., D.B, and J.A.G. Measurements and sample
preparation were done by S.G. and S.L. All other activities, including data interpretation,
conclusions, and manuscript writing, were carried out collaboratively by S.G, D.B. and J.A.G.
Additional Information
The authors declare no competing financial interests. Correspondence and requests for
materials should be addressed to J.A.G. ([email protected]).
Figure Legends
Figure 1. (a) The experimental device with dsDNA translocating through a graphene nanopore.
(b) Typical current blockades as dsDNA translocates through a 5 nm SiNx pore and through
graphene pores of various diameters (D). (c) dsDNA blocked currents through nanopores of
different diameters, defined by the open pore current. (d) Open pore currents through a series of
different diameter pores before addition of DNA. (e) Noise power spectral density of a 5 nm
graphene nanopore suspended across a 200 x 200 nm SiNx aperture (graphene 1) and a similar
13
nanopore across a 20 nm diameter SiNx aperture (graphene 2) with greatly reduced noise. A 5 nm
nanopore in SiNx with no graphene is also shown.
Figure 2. (a & b) Event plots for different diameter nanopores. Each circle corresponds to a
DNA translocation event. Blue, extended, unfolded molecule translocations (pH 10), green,
folded translocations (pH 10), and black, single stranded translocations (pH 12.5). Dashed lines
represent contours of constant ecd. (c & d) Single typical event current traces corresponding to
one of the events from the event plots in (a) and (b), for folded DNA, unfolded DNA and for
single stranded DNA. The event plots and current traces in panels (b) and (d) were all performed
with the same nanopore before the DNA was denatured (blue) and after the DNA was denatured
(black) by raising the solution pH to 12.5.
Figure 3. (a) Average current blockade for dsDNA translocations vs. nanopore diameter (solid
circles: graphene nanopores, open square: 30nm thick SiNX pores). Solid and dashed lines are fit
to the data using the numerical model. Inset: Nanopore conductivity vs. pore diameter. Red line
is simulated using best-fit values. (b) Calculated current density in a 2.5 nm graphene nanopore
traversed by dsDNA, the latter represented by a 1.93 nm cylinder. (c) Calculated Sensitivity
Gradient along a DNA molecule for the conditions in (b.
14
a
SiN pore
x
graphene pores
b
D = 5.0 nm
D = 4 nm
A
n
4
I0
DNA
c
B
I
)
A
n
(
I
15
10
7
6
5
4
3
6
10
14
18
Open pore current I (nA)
0
400 ms
d
25
20
)
A
15
n
(
0
I
10
5
2
4
6
D (nm)
pore
)
Z
H
/
2
A
n
(
D
S
P
-310
-510
D = 3.2 nm
D = 2.7 nm
e graphene 1
graphene 2
SiN
x
-710
-910
410
210
010
Frequency (Hz)
)
A
n
(
B
I
5
4
3
2
1
0
a
ds-DNA
5
4
3
2
1
)
A
n
(
B
I
D = 4.5 nm
pore
0
200
400
Event Duration (ms)
c
IB
IB
0
0
-1
-2
-3
)
A
n
(
B
I
b
ds-DNA
ss-DNA
D = 3.2 nm
pore
0
400
200
0
200
Event Duration (ms)
d
0
-1
-2
-3
)
A
n
(
B
I
dsDNA
IB
ssDNA
400
IB
400 ms
400 ms
2
a
)
A
n
(
B
I
4
3
2
1
0
2
3
4
5
6
Ion Conducting Diameter, D [nm]
IC
TEM Diameter, D [nm]
pore
4
3
5
6
7
4
150
)
A
100
n
(
G
50
2
4
6
D [nm]
pore
b
+160mV
graphene
0
1 nm
dsDNA
ground
-2
0
+2
D
i
s
t
a
n
c
e
[
n
m
]
0
8
6
2
4
Sensitivity
Gradient (nA/nm)
2
2
)
m
/
A
G
(
c
j
0
|
1002.3552 | 1 | 1002 | 2010-02-18T15:56:39 | Entropy analyses of spatiotemporal synchronizations in brain signals from patients with focal epilepsies | [
"physics.bio-ph",
"physics.data-an",
"q-bio.NC"
] | The electroencephalographic (EEG) data intracerebrally recorded from 20 epileptic humans with different brain origins of focal epilepsies or types of seizures, ages and sexes are investigated (nearly 700 million data). Multi channel univariate amplitude analyses are performed and it is shown that time dependent Shannon entropies can be used to predict focal epileptic seizure onsets in different epileptogenic brain zones of different patients. Formations or time evolutions of the synchronizations in the brain signals from epileptogenic or non epileptogenic areas of the patients in ictal interval or inter-ictal interval are further investigated employing spatial or temporal differences of the entropies. | physics.bio-ph | physics | Entropy analyses of spatiotemporal synchronizations in brain signals from patients with
focal epilepsies
Çağlar Tuncay
[email protected]
Abstract:
The electroencephalographic (EEG) data intracerebrally recorded from 20
epileptic humans with different brain origins of focal epilepsies or types of seizures, ages and
sexes are investigated (nearly 700 million data). Multi channel univariate amplitude analyses
are performed and it is shown that time dependent Shannon entropies can be used to predict
focal epileptic seizure onsets in different epileptogenic brain zones of different patients.
Formations or time evolutions of the synchronizations in the brain signals from epileptogenic
or non epileptogenic areas of the patients in ictal interval or inter-ictal interval are further
investigated employing spatial or temporal differences of the entropies.
Key words: Distribution, Shannon entropy, Interdependency, Synchronization, Pre-ictal
state, Post-ictal state, Seizure.
Pacs: ; 87.19.xm; 87.19.Nn; 87.19.le; 87.90.+y; 05.45.Tp; 05.90.+m
Introduction: Automatic (computer based) detection or prediction of epileptic seizures is a
challenging subject in quantified electroencephalography [1-13] where the underlying
dynamics for the recordings are usually not completely known. Thus, it is difficult to select a
priori most suitable method for an analysis [14].
Fortunately, the ictal state is characterized by occurrence of synchronous oscillations and
two scenarios of how a spontaneous seizure could evolve are suggested in [15-16]: If a seizure
is caused by a sudden and abrupt transition, then it would not be preceded by detectable
dynamical changes in EEG data. Such a scenario is suggested for the initiation of seizures in
primary generalized epilepsy. On the other hand, if the transition occurs gradually then it
could be detected, at least in principle. This type of transition is proposed to be more likely in
focal epilepsies. In this case, the seizure begins in a restricted brain region and either remains
localized or spreads to the adjacent cortex.
These hypotheses are considered and tested by various methods in several papers. For
example in [17 and 18], the interdependence between intracranial EEG electrodes in human
subjects with epilepsy is investigated. It is reported in both publications that there is strong
evidence for non-linear interdependence between the focus of epileptic activity and other
brain regions and this interdependence can be detected for up to 30 seconds prior to a seizure.
Different researchers [19] studied single electrode scalp EEG from 60 healthy humans and
found that non-linear structure is only present weakly and infrequently. More recently, the
same hypotheses were tested by comparing 30 measures in terms of their ability to distinguish
between the inter-ictal period and the pre-seizure period [20 and 21]. These publications
provide statistically significant evidence for the existence of a pre-ictal state which is
supported by numerous clinical evidence involving an increase in cerebral blood flow [22-23],
oxygen availability [24] and blood-oxygen-level-dependent signal [25] as well as changes in
heart rate [26-28] before seizure onsets. Thus, the interdependence between several brain sites
and the existence of pre-ictal state may be important for computer based investigations of
focal epilepsy.
The above mentioned approaches based on amplitude, time or frequency domain
measurements [21, 29-31] have a shared property in that they all try to figure out whether
- 1 -
there is any common information between the EEG time series as an indication of their
relationship. Then, direct investigations by means of information-theoretic approaches can be
favored [29]. Following this idea in this work, multi channel univariate time dependent
entropy analyses are performed on various electroencephalograms intracerebrally recorded
from humans with focal epilepsies [32]. The basic aim is to predict and detect numerous
seizure onsets covered in the investigated data. In the meantime, several properties in the EEG
data will be disclosed and used to distinguish between the inter-ictal period and the pre-
seizure period [20-21]. It is clear that a method capable of predicting the occurrence of
seizures from the electroencephalograms of epilepsy patients would open new therapeutic
possibilities.
Time dependent entropy:
In this study, time dependent interrelated multi channel EEG
activities of numerous humans are investigated. The ability of an information measure
(entropy) index [6, 33, 34], which is defined below, is determined to distinguish between pre-
ictal, ictal or post-ictal periods as well as the inter-ictal interval. The multi channel univariates
are cut into consecutive sections of convenient length of 10 seconds, and then entropy is
calculated for the distribution of data in each of the windows which do not overlap [35, 36]. In
other words, the analyses are performed in a time window of defined length and this time
window is moved step wise to produce a continuously updated entropy spectrum for the
recordings from different brain sites. Note that the model used is time invariant as stipulated
in [35].
Entropy is known to be a quantity describing the amount of disorder in a system or a
measure of uncertainty in the information content of a signal. Shannon entropy [37 and 38] is
used to investigate various properties of integrated neuronal activity in this work, where the
recorded voltages (Xk(i)) from the contact position (k=1-6) have integer values in micro Volt
(μV) [32]. Let Q(Xk(i)) be the frequency of Xk(i) in the data and be abbreviated as Qk(i). The
(discrete) empirical distribution Pk(i) is obtained by normalizing Qk(i);
where N is the number of samples (which is 2560 per window since the sampling rate is 256
Hertz (Hz) in [32] and the length of a time window is 10 seconds in this contribution). In what
follows, ED refers to empirical distribution and an ED is said to be discrete uniform (or
homogeneous) if P(i) ∝ 1/N or delta-like if data are distributed in a narrow peak around a
voltage.
Then estimation of Shannon entropy (Sk) of a time window of a univariate recorded from
the contact position (k) is
In Eq. (2), the summation is over the states (i) which are accessible with probability Pk(i)≤1
and log is logarithm with any base [37 and 38] which is taken as the natural logarithm, here.
Note that Sk is zero for delta-like ED. Moreover, Sk attains its maximum value which is
ln(N) [29] if the ED of a given set of N samples is homogeneous. Thus, an entropy measure
index (χk) may be defined as
Sk = -∑i=1;N Pk(i)log(Pk(i)) .
χk = Sk/ln(N)
Pk(i) = Qk(i)/∑i=1;N Qk(i) ,
(3)
(2)
(1)
- 2 -
which is zero for delta-like ED or 1 for homogeneous ED and 0<χk<1 otherwise. (See,
Appendix for further discussion of the subject.) One may refer to [20, 21, 39, 40] for several
utilizations of the parameter in Eq. (3) with different aims.
In this work, the ED for each time window starting at the time (τ-Δτ) and ending at the time
(τ) with Δτ=10 seconds are calculated and a continuously updated entropy index spectrum
χk(τ) is obtained for each univariate recorded from the contact position k=1-6.
The number of samples, N in Eq. (3), is the same for the windows and hence it is possible to
select these time windows with small values of χk in the time profile (χk(τ)) when the EEG
signal in the window is narrow (oscillations with small magnitudes about a voltage). On the
other hand, the value of χk is close to 1 when the EEG signals are broad (oscillations with
large magnitudes). Numerous further criteria may be applied to predict the characteristics of a
sample distribution, if needed [29, 39, 41-45].
Space and time differences of the entropy indices: Due to the occurrence of synchronous
EEG oscillations in the ictal state as discussed in the Introduction section, it is expected that
the time dependent indices (χk(τ)) become similar at times close to or during a seizure. In
other words, the values of the indices χk(τ) and χk’(τ) are synchronously close to each other
when the brain sites for the contact positions (k) and (k’) are coupled. Thus, the pairwise
differences Dk,k’(τ) of χk(τ) and χk’(τ)
may provide information about the strength (with small absolute value of Dk,k’(τ)) and
duration (number of the time windows with small absolute value of Dk,k’(τ)) of a coupling
between two brain sites. Note that Dk,k’(τ) may be considered as the approximate spatial
derivatives of the index profiles for those contacts which are closely positioned.
The seizure time terms last for several seconds or minutes and they are clearly short with
respect to several hours (h) for the registration periods of the data from a patient. Hence, a
new parameter (z) is used to help the inspection of the similarities in the time profiles of χk(τ):
where (.) stands for the absolute value. Obviously, z(τ) defined in Eq. (5) is big when the
values of the entropy indices (χk(τ)) are close to each other. Thus, the predictive power of the
marker depends on the strength of the couplings before or during onsets.
In this work, the first and second order successive time differences of the entropy indices,
χ’k(τ) and χ’’k(τ), respectively are also investigated with the aim of obtaining a better
description of the properties of integrated neuronal activity in ictal and inter-ictal intervals,
where
z(τ) = 1 / ∑k=1;6,k≠k’∑k’=1;6Dk,k’(τ)
Dk,k’(τ) = χk(τ) - χk’(τ)
(5)
(4)
χ’k(τ) = χk(τ) - χk(τ-Δτ)
(6)
χ’’k(τ) = χ’k(τ) - χ’k(τ-Δτ) .
and
In Eqs. (6) and (7), Δτ is equal to 10 seconds which is the length of a time window. These
time rates of change provide additional information for the formations and time evolutions of
the synchronized behaviors in entropy indices (χk(τ)). Thus, the attempts to understand the
underlying mechanisms of the synchronizations in the EEG signals of the patients entail the
analysis of the time series χ’k(τ) and χ’’k(τ) as well as χk(τ) and the spatial differences (χk(τ)-
(7)
- 3 -
χk’(τ)) in the presented work. Obviously, comparison of the distributions of these quantities is
further illuminating.
Correlations: Auto-correlation and cross-correlation functions are widely used in the
literature to investigate the interdependence between physiological signals [29-31]. However,
the values of the spatial differences of the electrodes (Dk,k’(τ)) or their distributions may be
used confidently with the same aim. For example, narrow and high peaks around zero in these
distributions will imply strong correlations and long durations, respectively between the
related individual recordings (Xk and Xk’) or vice versa. As a result, sample distributions of
the values of Dk,k’(τ) for several k or k’ are used to investigate the strengths and durations of
the correlations in this contribution.
Material:
Some brief information about all of the patients (PatNM) is given in Table I
where the patients are grouped with respect to their origins of epilepsy. It is declared in [32]
that “The EEG database contains invasive EEG recordings of 21 patients suffering from
medically intractable focal epilepsy.” Yet, the data from the Pat02 is missing there. Thus the
available data are from 20 patients. Let us follow the lines of the declaration: “In order to
obtain a high signal-to-noise ratio, fewer artifacts, and to record directly from focal areas,
intracranial grid-, strip, and depth-electrodes were utilized. The EEG data were acquired using
a Neurofile NT digital video EEG system with 128 channels, 256 Hz sampling rate, and a 16
bit analogue-to-digital converter. Notch or band pass filters have not been applied.
For each of the patients, there are data sets called ‘ictal’ and ‘inter-ictal’, the former
containing files with epileptic seizures and at least 50 minute pre-ictal data. The latter
containing approximately 24 h of EEG-recordings without seizure activity. At least 24 h of
continuous interictal recordings are available for 13 patients. For the remaining patients
interictal invasive EEG data consisting of less than 24 h were joined together, to end up with
at least 24 h per patient. For each patient, the recordings of three focal and three extra-focal
electrode contacts is available.”
This pool of data has been downloaded by nearly hundred research groups from different
countries [46] and treated with several aims [47-52].
The data recorded from epileptogenic zones will be designated with (k=1-3) and those from
non epileptogenic ones with (k=4-6) for each patient, here. As a result, data recorded from
each of the six contact positions (k) in the ictal interval or in the inter-ictal interval of each
patient are cut into consecutive sections of length of 10 seconds. These time windows include
2560 data per individual recordings (k) with 256 Hz sampling rate. Then, χk(τ) is calculated
for the distribution of the data in each of the non overlapping windows. Hence, six time
dependent trajectories of the entropy measure (χk(τ)) are obtained for the data recorded from
each patient in ictal interval and another 6 trajectories for the data from the same patient in the
inter-ictal interval. The results are compared for a characteristic behavior significant to pre-
ictal, ictal or inter-ictal states. The seizure terms of each patient are indicated in [32] and the
given numbers are utilized to select the seizure time windows in this work. The one-hour-
data-blocks with consecutive integer numbers are assumed to be continuous due to the
declaration [32].
Please note that numerous pin-like topographies are expected to occur temporarily in the
index profiles for the collective behaviors where the topographies will be upward if the
signals broaden in the meantime, or vice versa; namely, downward pin-like patterns will
appear if narrow signals accompany the synchronized interdependencies.
The original results are presented in the following section. The last section is devoted to
discussion and conclusion.
- 4 -
Results:
In this section the continuously updated entropy index spectrum for each of the
individual recordings from different brain sites (χk(τ) with k=1-6) of 20 patients with different
origins of epilepsy, seizure types, ages and genders are analyzed. The results are exemplified
for six patients with different origins of epilepsy, types of seizures, ages and sexes where
various characteristic behaviors are found in the data from pre-ictal, ictal or post ictal terms.
The available [32] electrode types or electrode contacts of the considered six patients are
shown in Figure 1. The theoretical results for χk(τ) or z(τ) are plotted with a linear axis on the
left and logarithmic axis on the right, respectively where the common horizontal axis shows
the time (τ) with the unit of 1 h, for 360 windows and the seizure terms are indicated by
arrows in Figures 2-5, 7, 9, 11, 13, 15, 18, 21, 22, 25. The number of the significant digits is
taken three after the decimal point in the results. Moreover, the bin sizes of the distributions
displayed in Figures 6, 8, 10, 12, 14, 16, 17, 19, 20, 23 are the same, which is 0.001 and that
in Figure 24 is 0.002. The parameter for k in the legends of the figures indicates the electrode
number.
CASE 1: Pat03 with Frontal Lobe Epilepsy
The time spectra of the entropy indices
(χk(τ)) with k=1-6 of the EEG data recorded from the Pat03 during the ictal interval and the
inter-ictal interval are illustrated in Figs. 2 (a) and (b), respectively. It may be inspected that
all of the (four out of four) indicated seizures [32] are detected by means of the marker z(τ) in
Fig. 2 (a). Thus, all of the seizures of the patient are accompanied by strong simultaneous
couplings (synchronous interdependency) between the brain sites for the contact positions.
Please note that the patterns of the trajectories of the entropy indices in the ictal interval and
in the inter-ictal interval are different: The profiles are more oscillatory in Fig. 2 (a) with
regard to those in Fig. 2 (b). However, several strong or weak couplings occur temporally in
both of the intervals as the big values in the time courses of (z(τ)) indicate in Figs. 3 (a) and
(b). Here, a distinguishing property is that the coupled entropy indices do not come close to 1
in the inter-ictal interval. Moreover, no formations similar to these which occur nearly 20
minutes before the imminent seizure onset in Fig. 3 (a) can be found in those profiles for the
inter-ictal interval; see, Figs. 2 (b) or 3 (b). Furthermore, similar pre-ictal formations occur
before the known seizures of the Pat03 as exemplified in Figs. 4 (a) and (b) where the data are
continuous for 2 h in each plot. It should be noted that the temporal couplings; namely, the
pin-like topographies covering a few windows are not strong during the seizure indicated in
Fig. 4 (b) with respect to those in Fig. 3 (a). However, the values of the indices are close to 1
with broad signals in all. Moreover, various pin-like formations occur also after the seizures
for nearly 20 minutes. Therefore, it may be claimed that the spontaneous occurrences of
individual strong or weak couplings are not decisive for the initiation of a seizure, where an
important criterion is the emergence of a pre-ictal stage as stressed in [20-21].
In summary, a pre-ictal stage originates nearly 20 minutes before the first indicated seizure
(Fig. 5 (a), which is the same as Fig. 3 (a) with different time domain) and it evolves
intermittently through several coupling epochs, each of which continues nearly one minute
covering five or six windows, usually. It is reported in [17 and 18] that, the interdependences
can be detected for up to 30 seconds prior to a seizure. Also, see [14-16], and [20] and [21].
The indices come very close to each other around 1 during the seizure terms. The transitions
from an ictal state to inter-ictal state also continue for nearly 20 minutes with various strong
or weak coupling terms as in the pre-ictal stage and the index values decrease after the offsets;
namely, in the post-ictal stage, to their log-run or inter-ictal averages. Thus, a cycle for a
seizure is completed. Similar cyclic behavior consisting of pre-ictal, ictal and post-ictal states
accompany all of the seizures of the patient (see, Figs. 4 (a) and (b) for example).
It is known that various focal seizures begin in a restricted brain region and either remain
localized or spread to the adjacent cortex, as considered in the Introduction section. Let us
- 5 -
focus on 3 (a) or 5 (a) to inspect the approach, onset and offset of the first seizure. The pre-
ictal stage is presumably originated (due to an unknown biological reason) at about τ=1.45 h
where the entire entropy index values decrease for a few minutes. This means that the in-
focus or out-focus contact positions are temporarily coupled. The couplings are weakened in
the next five or six minutes. Afterwards, the strength and occurrence frequency of the
couplings increase intermittently with time and the tops of the pin-like topographies ascend
towards 1. All of the contact sites are coupled at about τ=1.65 h, i.e., a few minutes before the
onset of the seizure which occurs at about τ=1.7 h; where the maximum values of the indices
with k=1-6 are nearly the same. During the seizure, integrated neural activity reaches a
climax, where all of the indices have nearly the same value which is very close to 1 and EEG
signals broaden. Apparently, more brain sites are coupled with increasing strengths while
approaching to the seizure onset. It may be claimed that the spread of the seizure is not a
continuous process but an intermittent one since strong couplings occur intermittently. The
relaxation of the indices is gradual and accompanied by numerous couplings, which are also
intermittent. Figs. 5 (b) and (c) show the first and second order successive time differences
(approximate time derivatives) of the indices with the same time domain of Fig. 5 (a), where
the intermittent couplings are clear. Therefore, not only the indices but also their time
differences are coupled at several time epochs. Moreover, the brain electricity may be unique
and locally modulated when close to or during a seizure. (Various different suggestions on the
subject may be found on page 198 in [39].) This subject will be held while investigating other
cases for different patients and further discussed in the last section.
Let us now treat the distributions of the index values for the data from the Pat03 in the ictal
interval (Fig. 6 (a)) or in the inter-ictal interval (Fig. 6 (b)), where the aim is to search for
statistical evidence about the electrical couplings between different brain sites. The plots for
the data from the inter-ictal interval are evidently parabolic in logarithmic scale (or Gaussian
in linear scale) whereas those for the ictal interval depict important deviations from Gauss,
where the deviations are outstanding especially for the data from the out-focus electrodes,
with k=4-6. This shows that the out-focus sites are strongly affected by the epileptogenic
activities.
Note that the modes of the ED plots of the index values occur at the intermediate values
which match the values for the pre-ictal and post-ictal stages. Hence, these stages can be
predicted if the modes of the ED plots of the index values show certain deviations from
Gauss. Conclusively, the ED plots of the entropy index values can be utilized to distinguish
between the inter-ictal data and ictal data or more specifically, the inter-ictal period and the
pre-ictal period (see, [20-21]).
The time series of the mutual arithmetical differences of the in-focus (k=1-3) or out-focus
(k=4-6) index values are displayed in Figs. 7 (a) and (b), respectively which cover the first
seizure. Note that the differences remain close to each other around zero for a few minutes
after the seizure offsets and depart gradually afterwards. This behavior can also be monitored
during the other seizures of the Pat03 (Fig. 7 (c)) and the seizures of several other patients as
will be considered in further cases.
The over-all behavior of these electrode differences can be understood in terms of the
distributions of their values as displayed in Figs. 8 (a) and (b). Inspection of these figures
shows that the in-focus and out-focus electrodes are more strongly correlated in the ictal-
interval than in the inter-ictal interval since the modes occur at zero in Fig. 8 (a). Moreover,
the heights of these modes are nearly the same for all of the electrodes which indicate a strong
collective behavior where the heights are much bigger than the number of the windows for the
seizures. This is because numerous very strong couplings occur in the pre-ictal and post-ictal
stages as well as in the ictal stages. However, the index values come out very close to 1 for
these strong couplings only during the seizures, which is a distinguishing property for the
- 6 -
onsets. Furthermore, all of the out-focus electrodes and only two of the three in-focus
electrodes (k=1 and k=3); namely, FE1 and FD1 (Table I and Fig. 1) are strongly correlated in
the inter-ictal interval since the highest and narrowest peak around zero comes out for these
electrode differences (DFE1,FD1) in Fig. 8 (b). Hence the total duration of the DFE1,FD1 is the
longest in-between the in-focus or out-focus couplings during the inter-ictal interval.
Whereas, all of the out-focus electrodes are strongly coupled in both of the intervals where the
couplings are stronger in the ictal interval, Figs. 8 (a) (also, Fig. 7 (c)). Note that, the plots for
D1,2(τ) and D2,3(τ) are nearly anti-symmetric about zero. This is because, the time profile of
χ1(τ) is similar to that of χ3(τ); namely, (χ1(τ)↔χ3(τ)) where the symbol (a↔b) is used to
designate the similarity (or coupling) between a and b. Two trajectories are said to be similar
if their time courses are the same up to the minor fluctuations, in this work. Hence, the ED
plot for D1,3(τ) depict a narrow peak around zero since D1,3(τ)=χ1(τ)-χ3(τ)↔0 due to Eq. (4).
Therefore, (χ1(τ)-χ2(τ)) is similar to (χ3(τ)-χ2(τ)) and D1,2(τ)↔D3,2(τ)=-D2,3(τ). Moreover, the
in-focus electrode k=1 is strongly coupled with all of the out-focus electrodes in the ictal
interval whereas the in-focus electrode k=3 is strongly coupled with all of the out-focus
electrodes in the inter-ictal interval (not shown). Hence, the strength or durations of the
synchronous electrode interdependence may vary with the time.
The distributions of the values of the first or second order temporal differences are (Eqs. (6)
or (7), respectively) exponential and symmetric about zero for each electrode in the ictal
interval or in the inter-ictal interval (not shown). This subject will be discussed in detail in
further sections.
CASE 2: Pat14 with Fronto/temporal Epilepsy
Only three of four indicated seizures are
detected in this case (Fig. 9 (a)) where the first indicated seizure which occurs a few minutes
after an abrupt decrease in the indices (for a narrow signal) is missed (Fig. 9 (b)). During this
or other downward pin-like behaviors in the ictal and the inter-ictal data, the electrodes show
intermediately strong cross-correlations as displayed in Fig. 9 (c). Moreover, a diversity of
EEG signals from narrow to broad signals is generated at different times in this case. (Note
that there are no downward pin-like patterns for narrow signals in the previous case.)
The entropy profiles from the in-focus electrodes have big magnitudes and fluctuate with
small magnitudes around nearly 0.9 whereas these from the out-focus positions fluctuate with
big magnitudes around nearly 0.85 in the ictal interval (Figs. 9 (a) and (b)). The values of the
oscillations are above 0.8 also in the inter-ictal interval aside from the frequently occurring
downward pin-like behaviors (see Fig. 9 (c), for a short time domain).
Durations of the pre-ictal and post-ictal states are nearly 30 minutes or longer, as can be
seen in Fig. 9 (c) where the data are continuous for 3 h and two ictal cycles consisting of pre-
ictal, ictal and post-ictal stages are included. The profiles are inter-ictal-like up to τ∼4.6 h
where weak couplings occur between the electrodes and the entropies little increase. The
following pre-ictal state for the onset at τ∼4.8 h involves several weak synchronizations and
these interdependencies evolve into strong ones during the seizure which starts and terminates
abruptly. The post-ictal state lasts for nearly 30 minutes with numerous strong and
synchronized interdependencies. The adjacent inter-ictal state also lasts for about 30 minutes,
up to τ∼6.0 h while strong couplings start which become stronger at τ∼6.2 h. Thus, the pre-
ictal state for the next seizure which occurs at τ∼6.6 h initiates. Note that the values of the
entropy indices are bigger in the considered post-ictal stages with respect to these in the inter-
ictal or pre-ictal stages.
The first and second order time differences of the individual indices also display the
couplings (upward or downward pin-like topographies) in the ictal interval and the inter-ictal
interval (not shown). Moreover, the ED plots of the values of these differences are symmetric
and exponentially decreasing around zero as in the previous case for the Pat03 (not shown).
- 7 -
On the other hand, the ED of the ictal and inter-ictal index values are characteristically
different (as in the previous case) with some noisy behaviors in both of the ED (in this case),
as can be seen in Figs. 10 (a) and (b). The noisy behaviors and deformations from Gauss in
Fig. 10 (a) are due to the pre-ictal and post-ictal states as discussed in the previous case
whereas these in the data from the inter-ictal interval can be attributed to the time lengths of
the patient’s pre-ictal or post-ictal stages, both of which are nearly 30 minutes or longer.
Thus, various data from the pre-ictal or post-ictal epochs might have been mistakenly counted
within the one-hour-data-segments of the inter-ictal EEG in [32]. Namely, various pre-ictal
states manifest weak or intermediately strong couplings without leading to completed seizure
onsets, in the time epochs far from the experienced seizures and these data are counted for the
inter-ictal interval. However, the ictal and inter-ictal intervals can be clearly distinguished in
terms of the ED plots since these plots, especially for the out-focus electrodes from the ictal
interval show important deformations with respect to those from the inter-ictal interval as in
the previous case.
Moreover, the maximum values of the indices of this patient are very close to 1 (see, Fig. 10
(a) and (b) also Figs. 6 (a) and (b), 14 (a) and (b), 23 (a) and (b)) whereas the minimum values
are 0 due to various narrow signals. Hence very broad and narrow brain signals are generated
in this case.
Various simultaneous couplings between the local electrodes of the Pat14 can be utilized to
detect seizure cycles as considered in the previous case. Yet, the focus will be on the
simultaneous couplings between the non-local electrodes, here; see, Figs. 11 (a) and (b).
Although the cyclic behavior is not clear, the ability of the electrode differences to capture the
seizure term is obvious in Fig. 11 (a). Moreover, the differences D1,5(τ) and D2,5(τ) (the dotted
lines in different colors in Fig. 11 (a)) follow each other closely in the given time domain.
This feature implies that D1,2(τ)↔0 due to Eq. (4); namely, the contact positions at k=1 and
k=2
are
strongly
correlated: χ1(τ)↔χ2(τ)
thus,
(χ1(τ)-χ5(τ))↔(χ2(τ)-χ5(τ))
and
D1,5(τ)↔D2,5(τ).
The related cyclic behavior can be inspected in Fig. 11 (b) which shows the time courses of
a different set of non-local electrode couplings. The first and second integrated behaviors in
the profiles towards 0 at τ∼3.2 h and τ∼3.3 h, respectively indicate the pre-ictal stage. These
couplings are very strong during the seizure and weakened afterwards. They continue for
several minutes in the post-ictal stages as can be inspected in Fig. 11 (c) for different seizure
terms, where all of the time profiles of the non-local couplings are designated by dotted lines.
The ED plots in Fig. 12 (a) and (b) show that several simultaneous non-local cross-
correlations are weak but they follow each other closely through the ictal interval possibly
with a time phase difference; see, also Fig. 11 (a). (Please note that the frequency or time
phase domain analyses are kept beyond the scope of this contribution.) Figs. 12 (c) and (d) are
for the cross-correlations of local electrodes which depict that the cross-correlations may vary
with time. For example, the D1,3 follows the D2,3 or D1,2 in different time windows and with
different magnitudes.
Relying upon the predictions made in the previous paragraph, it may be remarked that the
unique brain electricity may be modulated differently at different times in different brain sites,
but in coordination with some other site(s). Because, the partnerships in the brain couplings
continue for long times yet the partners may be exchanged temporarily.
CASE 3: Pat20 with Tempo/Parietal Epilepsy
The data registration from the Pat20 in the
ictal interval (the seventh one-hour-data-block; 020322aa_0054_k with k=1-6 in [32]) is
interrupted (at i=837632 with k=1-6) for several minutes (Fig. 13 (a) at τ∼6.9 h) and no
reason for the disconnection is given in [32]. The brain states might have been affected
radically if some medical treatment were performed during the disconnection. For the effects
- 8 -
of medications such as drugs and anesthetics such as nitrous oxide on the brain states, or
several examples of the related Shannon entropy applications, see [53-56] and [36],
respectively and the references therein. Whatsoever this situation is, three out of four
indicated seizures are detected in this case. Yet, numerous false alerts are produced due to the
synchronized behavior in the data except those from the electrode k=4 as exemplified in Fig.
13 (b) where the data are continuous for 3 h. In the same figure, the pre-ictal stage for the first
seizure, which is undetected, is very long and consisting of several sub-stages with various
durations. At τ∼0.2 h, the index values decrease collectively. Afterwards, the values increase
and various couplings occur between the contact positions except at k=4. These oscillations
continue till τ∼0.7 h where a new collective decrement similar to the previous one occurs.
This time, the separation between the profiles, from the contact position at k=4 and others, are
obvious. The tops of the pin-like topographies increase gradually and approach 1 very closely
at τ∼1.6 h where a seizure might have been expected to onset as z(τ) indicates. However, the
values decrease abruptly and this state continues for nearly 20 minutes till τ∼1.9 h where an
onset is realized. Therefore, numerous false attempts for an onset can be found in both of the
ictal or inter-ictal data from this patient. (See, Fig. 3 (a) for two similar incomplete formations
before the experienced seizure of a different patient.)
It can be observed in Fig. 13 (b) and (c) that the onset and offset of the seizure at τ∼1.9 h are
abrupt. Moreover, the electrode positions at k=1 and k=6 or k=2-5 are coupled strongly right
after the offset for nearly 5 minutes as depicted in Fig. 13 (c). The electrodes change couples
afterwards and this epoch continues for nearly 20 minutes. Various false inclinations for a
seizure onset might have been observed in the several epochs of the ictal data from this
patient (for example, between τ∼2.6 h and τ∼2.9 h).
The couple exchanges can be monitored in the ED plots of the index values from the ictal
interval and the inter-ictal interval (Figs. 14 (a) and (b)) where the collective phenomena are
clear for big values of the indices from both of the intervals. Note the strong coupling between
the electrodes k=1 and k=6 where this partnership continues throughout the ictal interval and
inter-ictal interval.
The time series for the time differences also exhibit the pin-like topographies for the
couplings and the ED plots of their values come out symmetric and exponentially decreasing
around zero. The local or non-local spatial couplings may run with the same partners in
several epochs and the couples may be changed in others. Moreover, these differences may be
utilized to detect the same seizure onsets which are captured by means of the marker. Thus,
further investigations of the spatiotemporal differences of the electrodes do not provide new
information in this case, because they are obtainable using different methods as shown in the
previous analyses.
However, it should be noted that the aforementioned (the first paragraph in this CASE)
disconnection occurs at τ∼6.9 h but the index values decrease to zero much earlier than the
disconnection; namely, at τ∼6.4 h and this behavior continues till the disconnection. The
index profiles with k=1-6 decrease to zero much earlier than the disconnections occur also in
the data from several patients in different cases. Hence, there may be a causal relation
between the narrowness of the EEG signals before the disconnections and the reason(s) for
the disconnections. This subject is further discussed in the last section.
CASE 4: Pat11 with Parietal Epilepsy
The entropy indices from the Pat11 are obviously
non-stationary as displayed in Fig 15 (a). The indices from the continuous first two one-hour-
data-blocks have small values and the seizure term at τ∼1.05 h is undetected by the current
method. However, both of the two indicated seizures in the continuous data for the following
4 hours (2 h<τ≤6 h) are captured. Yet, the last seizure which occurs in the last continuous data
- 9 -
set (6 h<τ≤8 h) is missed. As a result, two of the four indicated seizures are detected in this
case.
It should be noted that the patterns of the profiles for the second and third seizure terms are
similar to each other and thus there are three sets of continuous data and three types of seizure
profiles in this case (Figs. 15 (a)-(c)). The first set of the continuous data deliver nearly
homogenous index profiles where no pre-ictal, ictal or post-ictal stages are distinguished.
Obviously, the first seizure does not show a cyclic behavior. Whereas, the cyclic
interdependence is clear in the second and third seizure terms (Fig. 15 (b)).
The index values gradually increase from τ∼2.0 h up to τ∼2.5 h and abruptly decrease to the
previous levels a little later (at τ∼3.0 h). Afterwards, they increase again gradually up to τ∼3.1
h and then relax. Both of these eras are clearly not inter-ictal-like but they resemble the pre-
ictal states for the imminent two seizures. Thus, they can be taken as failed alerts for a
seizure. A clear pre-ictal state for the onset at τ∼3.6 h stars at τ∼3.3 h where the out-focus
index values rise more rapidly than the in-focus ones. This behavior can be inspected also in
the pre-ictal stage of the onset at τ∼4.8 h. The onsets and offsets of both of these seizures are
abrupt and the couplings are not strong in-between them. The intermediate (inter-ictal) stages
are depicted at τ∼4.0 h and τ∼6.5 h both of which continue for nearly half an hour.
The ED plots of the ictal or inter-ictal indices are depicted in Figs. 16 (a) and (b),
respectively where the integration is clear for big values. However, the maximum values
come out bigger in the ictal interval (for the onsets). Moreover, the plots for the in-focus
electrode at k=1 and out-focus one at k=6 are similar in both of the intervals, which shows
that the k=1 and k=6 sites are less affected by the other sites. Whereas the same plots for the
data from the k=4 position show big variation from one interval to the other and hence it can
be claimed that the brain site at k=4 is more effected by the others at the times close to the
seizures.
The analyses of the ED plots for the spatial differences of the electrodes show that the in-
focus electrodes are strongly correlated in both of the ictal interval and inter-ictal interval,
Figs. 17 (a) and (b). Three further strong local or mixed cross-correlations come out in the
inter-ictal interval; between the positions at k=4 and k=6, or k=1 and k=6, and k=2 and k=6,
respectively. But, the non-local coupling between the sites at k=4 and k=6 weakens in the
ictal interval. Please note that numerous time windows involving disconnections or very
narrow signals (with zero index values) in the inter-ictal interval are omitted in Fig. 17 (b).
CASE 5: Pat15 with Temporal Epilepsy
The first four one-hour ictal data blocks with
k=1-6 are continuous which cover two seizures as depicted in Figs. 18 (a) and (b). The pre-
ictal state for the first onset starts at τ∼0.2 h and continues till τ∼0.9 h where the onset is
experienced. The post-ictal state is not clear for this seizure. As a result, the pre-ictal state for
the second seizure can be hardly detected. The second onset occurs at τ∼3.1 h and the post-
ictal stage of this seizure continues till τ∼3.7 h without a doubt.
Also the next three one-hour-data blocks with k=1,6 are continuous and they cover one
seizure (Figs. 18 (a) and (c)) where the indices oscillate with big magnitudes. The oscillations
are minor in the time interval between τ∼4.8 h and τ∼5.2 h which may be taken as the
intermediate state. The pre-ictal state starts at τ∼5.2 h as distinguished in terms of collective
behavior. In other words, the pin-like synchronous interdependencies start at that time and
gain importance with time; namely, the maximum values ascend towards 1 and the frequency
of the occurrences increase with time. At τ∼5.2 h, the integrated behavior reaches a climax,
where all of the indices have nearly the same value which is very close to 1. Moreover, the
indices drop two times; one during the seizure and one right after the seizure offset. The index
values increase abruptly or gradually afterwards till τ∼5.85 h and this instant can be taken as
- 10 -
the termination of the post-ictal stage which indicates the completion of a seizure cycle. Note
that, the electrode couplings continue till τ∼6.05 h with big index values which become close
to 1 at τ∼6.3 h. The time epoch between τ∼5.85 h and τ∼6.05 h should be considered not as a
part of the last post-ictal term but a new pre-ictal state for an unrealized onset at τ∼6.3 h. Two
more similar inclinations which failed before an onset can be monitored in Fig. 18 (c). The
related clinical evidence is not reported in [32].
The last seizure onset occurs at τ∼8.6 h in the interval 7.0 h<τ≤10.0 h which contains
continuous data. The index values collectively increase towards 1 with various strength of the
couplings several times before the onset; such as at τ∼7.3 h and τ∼7.9 h. Yet, the strong
couplings start at τ∼8.3 h which can be taken as the beginning of the pre-ictal stage of the
seizure and thus, this stage continues for nearly 20 minutes. An abrupt decrease in the index
values occur within the seizure term and these decrements are followed by gradual
increments. Hence, the seizure offsets before the completion of the gradual increment term.
The values come close to each other around 0.95, that is, a few minutes after the offset and
then the post-ictal state occurs and continues till τ∼7.3 h.
All of the indicated seizures (four out of four) are detected in this case besides numerous
pre-ictal stages of the assumed uncompleted seizure cycles. The pre-ictal or post-ictal
transition terms of the realized seizures may last half an hour or longer. Thus, also the ED
plots from the inter-ictal interval come out with noise and various deformations from Gauss in
the range for the intermediate index values, as those from the ictal-interval in the declared
range of the index values; see Figs. 19 (a) and (b). Moreover, the tails of the plots for the big
values in Fig. 19 (b) indicate that, the inter-ictal data involve several pre-ictal stages for
uncompleted seizure cycles which are manifest also in the time profiles (not shown).
Furthermore, the couplings between all of the electrodes, with the exception of the one at k=4
are strong during the ictal-interval with long durations.
The individual behavior of the electrode at k=4 is clear in the ED plots for the spatial
differences from ictal interval and inter-ictal interval in Figs. 20 (a) and (b), respectively. Note
that no strong cross-correlations involve data from the site at k=4 in both of the intervals.
Moreover, the groups of the ED plots for D4,5 and D4,6 (local) or D1,4, D2,4 and D3,4 (non-local)
are similar to each other and the groups of the plots are nearly anti-symmetric about zero in
the ictal interval (Fig. 20 (a)). This anti-symmetry is exemplified in Fig. 21 (a) within the
three-hour-continuous data which cover the third seizure term as in Fig. 18 (c). This anti-
symmetry occurs since the plots of D1,4, D2,4, D3,4, and -D4,5 (=D5,4) and -D4,6 (=D6,4) come out
similar. In other words, the index from the electrode at k=4 follow a different trajectory than
the other electrodes where the differences in the trajectories of the indices from the electrodes
with k≠4 can be neglected (the central group in Fig. 20 (a)). As a result, the plots of Dk,4 with
k≠4 come out similar. Therefore, all of the brain sites except at k=4 couple strongly in the
ictal interval of this patient.
Fig. 21 (b) is the same as Fig. 21 (a) for a shorter time range and for all of the electrode
couplings, where the following properties are inspected: All of the electrodes are coupled
strongly for a while which implies that the brain electricity is unified at every site in these
time terms. Moreover, the seizure offsets whilst the electricity is unique. The couplings
become strong once again at τ∼6.8 h with a longer duration. Fig. 21 (c) covers the last seizure
where a similar behavior to the one described in the last lines of the previous paragraph is
manifest yet the offset occurs during the strong couplings of the brain sites.
The time profiles of the successive time differences of the individual recordings display the
temporal couplings and their values are distributed symmetrically and exponentially around
zero as in the previous cases (not shown).
- 11 -
CASE 6: Pat09 with Temporo/Occipital Epilepsy The continuously updated index values of
the patient oscillate with big magnitudes in ictal interval (Fig. 22 (a)) where, four out of five
of the indicated seizures are detected and several false alerts are obtained. The beginning of
the pre-ictal stage for the first seizure (τ∼1.05 h) is not clear. Weak couplings with short
durations are inspected at the beginning of the first one-hour-data-blocks with k=1-6. These
weakly synchronized interdependencies become stronger and the event frequencies increase
with the time. The post-ictal stage continues nearly 20 minutes after the offset and several
pin-like formations appear afterwards till τ∼2.0 h, which is the end of the first set of
continuous one-hour-data-blocks. The data for 2 h<τ≤6 h are also continuous and they involve
two seizures (Fig. 22 (a)) which are accompanied by clear pre-ictal states with the time
durations of nearly 30 minutes. These pre-ictal states can be distinguished from the
intermediate-states in terms of the increasing number of strong couplings with time. Whereas,
the time durations of the post-ictal states of these seizures are shorter than these of the pre-
ictal ones.
The most interesting feature in the results from the patient is the strong couplings between
all of the electrodes at 8.1 h<τ<8.4 h which continue for nearly 20 minutes (Figs. 22 (a) and
(b)). Nearly at the beginning of the eighth one-hour-data-blocks (010906ea_0070_k with k=1-
6), the entropies of the individual recordings follow each other very closely for about nearly
20 minutes yet no seizure is indicated in [32] for the current interval. Note that a similar
formation occurs also in the data from the inter-ictal interval as displayed in Fig. 22 (c), where
the data are continuous along the interval.
The ED plots of the index values from the ictal interval (Fig. 23 (a)) and inter-ictal interval
(Fig. 23 (b)) are capable of distinguishing these intervals, clearly. Note that the electrodes do
not exchange partners in the regime for small index values and the out focus electrode at k=6
is coupled with the in-focus electrodes in both of the intervals. The maximum values of the
indices are very close to 1 with very broad or homogeneous signals as in CASE 2.
The cross-correlations between the electrodes at k=5 and k=6 are strong in the ictal and
inter-ictal interval where the ED plots for these correlations and that for the group of D1,5,
D2,4, D3,4 and D3,5 are anti-symmetric. The temporal differences of the individual recordings
do not provide additional information since they are similar to those from the other patients.
These two subjects are further discussed in the next section.
In summary, nearly 700 million ictal or inter-ictal EEG data intracerebrally recorded from
20 patients with different ages, genders, origins of the epilepsy, seizure types and electrode
types or insertion places are investigated in terms of the continuously updated entropy indices
χk(τ) with k=1-6. A common property predicted in the results, which are exemplified in detail
for 6 patients, is that the index values increase and then decrease several times before the
imminent seizures and reach at locally maximum values close to 1 during the onsets. The
characteristic features of these similarities are: 1) The indices from the ictal interval may
fluctuate about different values for different k with weak couplings. This era may be called
the intermediate (inter-ictal) state during which no seizure is expected to occur soon. 2) When
the number of the couplings per a given time interval and their strength increase with the time,
this means that the patient is approaching an onset of a seizure. This era stars 20-30 minutes
before the onset and it may be called the pre-ictal state. If the tops; namely, the local
maximum values of the entropy indices increase and the synchronizations cover more
univariates from different brain sites then the onset is expected to occur within a few minutes.
3) The detected seizure onsets are accompanied by synchronized and strong couplings in the
entropy indices with values very close to 1. This means that the brain signals are unified at
several sites and broad during these seizures. 4) The indices relax after the offsets gradually in
most of the cases or abruptly in forms downward pin-like formations in a few cases; as in
- 12 -
Case 5, for example). This era may be taken as the post-ictal stage which continues for 20-30
minutes usually.
It may be remarked that the synchronization in the EEG signals can be monitored as a
parallelization in the time profiles of the entropy indices or their successive time differences,
χ’k(τ) or χ’’k(τ). The time courses of variation of these parallel behaviors are investigated also
in terms of the electrode couplings in order to better understand the underlying mechanisms.
Discussion and conclusion: The presented approach provides a reasonable description of
various spatiotemporal similarities in between the brain states before, during and after the
captured seizures. These similarities are predicted to be cyclic spatiotemporal synchronization
(coupling) of the brain sites where the spatial couplings may be local; namely, between two
(or three) in-focus electrodes or between two (or three) out-focus ones, or non-local; namely,
between numerous in-focus and out-focus electrodes. Moreover, the durations of these
partnerships or the partner sites may change with time.
The patterns of the integrated neuronal behavior observed in the entropy indices, as well as
in the spatiotemporal differences (approximate derivatives) of the indices, come out patient-
specific, which may be related to psychological fluctuations such as, vigilance (wakefulness,
drowsiness or sleep), attention and anxiety in a patient. These interrelationships are found to
be weak and infrequent in the inter-ictal intervals and when they intermittently evolve into
strong and frequent ones then this stage is distinguished as a pre-ictal state. Thus, the
epileptogenic processes are predicted to be intermittent. Similar results are reported in [58]
where long-term intracranial recordings of five patients by employing a measure of phase
synchronization are investigated.
Furthermore, it is observed that extremely strong couplings occur between the brain sites
during the predicted seizure onsets and the transitions from these onsets to post-ictal or inter-
ictal stages also follow patient-specific patterns in the time profiles of the entropy indices and
their spatiotemporal differences. As a result, all of the patterns in these time profiles are
clearly patient dependent. Therefore, not only the epileptogenic processes but also the
relaxation processes are patient-specific, intermittent and cyclic where a cycle for the entire
ictal process consists of three stages: (pre-ictal) preparation for, (ictal) realization of and
(post-ictal) recovery from a seizure.
However, the distributions of the values of the first order successive time derivatives of the
indices (Eq. (6)) seem patient independent as displayed in Fig. 24 where all of the available
data in [32] from 20 patients in ictal interval or inter-ictal interval are used. (The total number
of the utilized time windows with no disconnection or non-zero entropy values in Fig. 24 is
234,879.) It should be noted that the plots of these distributions are symmetric and
exponentially decreasing for each of the patients in ictal interval or inter-ictal interval as
considered in the previous section for various cases. Moreover, the ED plots of the values of
the second order successive time derivatives (Eq. (7)) come out similar to those in Fig. 24. It
can be claimed that the distribution curves of χ’k(τ) and χk’’(τ) with k=1-6 are similar not
only for focal epilepsies but for all of the epileptic and possibly healthy brains. (The tails, for
the big magnitudes, will come out presumably clear for healthy brains).
Furthermore, the presented approach can be considered successful since nineteen out of the
twenty six studied seizures are captured as discussed in detail in the previous section. This
constitutes a more than 70 % success rate. A similar success rate is achieved also in [58].
The types or insertion places of the used in-focus or out-focus electrodes are different (Fig.
1). Hence, the capability of the marker z(τ) (Eq. (4)) to capture the seizure onsets may change
from one case to the other. Moreover, a smoothed version of z(τ) (namely, the application of a
known smoothing function on z(τ)) or averages of z(τ) over a fixed number of the backward
- 13 -
or forward time windows at a time τ, may yield more precise detections where some big
powers of χk(τ) or χk’(τ) can also be utilized in Eq. (4).
It can be expected that the accuracy of the results may depend on the length of the time
windows used due to the non-stationary character of the epileptic EEG signals.
Length dependence: Various analyses of magnitude square coherence and phase coherence,
entropy or entropy and complexity show that the results depend on the data length; see Table
1 in [59] and the results in [60] or [61], respectively.
In the results of this work presented above, the window length is 10 seconds. Thus each
window involves 2560 data per electrode (k), since the sampling rate is 256 Hz. Therefore,
numerous bins in the ED of the EEG voltages (Qk(i) in Eq. (1)) come out empty (Qk(i)=0)
since the bin size is 1 μV and the range of the voltages is from -10,000 μV to 10,000 μV or
broader. Hence, the window length may be crucial for the results, especially when a spiky
behavior or narrow or broad signal is dominant in a given epoch. In this case, longer time
windows can be used in order to reduce the effect of these irregularities on the results and to
avoid abnormal values for the indices.
The values of the entropy indices decrease to zero before various indicated disconnections
occur during the data registration in [32]. For example, in CASE 3 in the Results section, the
index values of the Pat20 in the ictal interval descend towards zero at τ∼6.4 h (Fig. 13 (a)) and
the values remain zero till a disconnection occurs in the data registration nearly 30 minutes
later; namely, at τ∼6.9 h (i=837632 with k=1-6). This appearance is observed in the EEG
from the same patient in the inter-ictal interval or from different patients in both of the
intervals. Hence, there may be a causal relationship between several irregular behaviors in the
EEG signals and the simultaneous clinical symptoms of the relevant patients before the
disconnections. These symptoms are not reported in [32]. Moreover, the effect of these
irregular data may be reduced in the entropy estimations when longer time windows are used
as exemplified in Fig. 25(b).
Figs. 25 (a)-(c) are obtained using time windows of length 100 seconds in the presented
approach. In Figs. 25 (a) and (b), the index values come out smaller than those in Fig. 13 (a)-
(c) by a factor of nearly 1.3 (∼10.15/7.85=∼ln(25600)/ln(2560), see Eq. (3)). The time domain
of Fig. 25 (a) is the same as that of Fig. 13 (b) where the following three features can be
clearly observed: the ictal cycle of the experienced onset at τ∼1.9 h, several unsuccessful
inclinations for a seizure before the onset and the individual behavior of χ4(τ) (thick line in
dark cyan in Fig. 25 (a)). The individual behavior of χ4(τ) can also be observed in Fig. 25 (b)
where the data contains a disconnection as discussed in the previous paragraph and Results
section (CASE 3: Pat20). It must be noted that the index values are non-zero also after τ∼6.4 h
till the occurrence of the disconnection at τ∼6.9 h in Fig. 25 (b).
The successive differences of the indices (χk’(τ)) with the newly used longer time windows
(Δτ=100 seconds in Eq. (6)) are displayed in Fig. 25 (c) where the time domain includes the
first indicated seizure of the Pat20, at τ∼1.8 h which was undetected previously.
The patterns obviously depend on the length of the time windows; however, the
characteristic features of the results (for example, the cyclic topographies of the ictal patterns)
remain invariant within reasonable limits for the length of the time windows.
Couplings and brain electricity:
If the entropy indices from the brain sites at k and k’ are
similar (χk(τ)↔χk’(τ)) in a time interval then these sites are called coupled or synchronously
interdependent in this work as referred to above. Thus, the four fundamental states of an
epileptic brain; namely, the inter-ictal, pre-ictal, ictal and post-ictal states can be distinguished
in terms of the simultaneous interrelationships. The couplings are found to be infrequent and
- 14 -
weak in an inter-ictal stage, as in [19] for healthy brains. The event frequency and strength of
the couplings usually increase with time in a pre-ictal state [20 and 21] and they decline after
the offsets for a post-ictal stage. In the meantime the strengths reach a climax for a few
minutes for a seizure. If the number of the coupled brain sites also increase in a pre-ictal state
then the brain electricity and possibly the seizure may be taken as spreading.
Various similarities between the brain electricity recorded from different sites are detected
in this contribution for different cases and thus the spread of the brain electricity is discussed.
However, these results could not be associated with the seizure types because no further
information about the individual seizure types is available in [32]. In other words, the
predictive power of the presented approach depends on the number and strength of the brain
couplings at a time. Thus, if a seizure onset occurs either with weak or a few brain couplings
then it could not be detected. It is not known which of the noted seizure types [32]; namely,
simple partial (SP), complex partial (CP) or generalized tonic-clonic (GTC) onsets fits which
of the investigated seizures. The presented approach might provide statistical evidence
between a seizure type and brain couplings if individual seizure types were known.
Let us consider the aforementioned (two paragraphs above) similarity between the two
entropy index trajectories, χk(τ) and χk’(τ). It is clear that, if (χk(τ)↔χk’(τ)) then (χk(τ)-
χk’’(τ))↔(χk’(τ)-χk’’(τ)). Thus, Dk,k’(τ)↔0 and Dk,k’’(τ)↔Dk’,k’’(τ). Such a situation was
considered in CASE 1 for the Pat03 in the Results section where D1,3(τ)↔0 and D1,2(τ)↔
-D2,3(τ)=D3,2(τ) for 1 h<τ< 2 h which covers a seizure with unknown type, (Fig. 7 (c)).
Another example is given in CASE 2 for the Pat14 as depicted in Fig. 11 (a) where
D1,5(τ)↔D2,5(τ) (dotted lines) as a result D1,2(τ)↔0 as shown in Fig 12 (a).
Thus, if two brain sites k and k’ are coupled strongly then Dk,k’’(τ) and Dk’,k’’(τ) with
k’’≠k’≠k, follow each other very closely. The stronger is the coupling between the brain sites
at k and k’, the smaller is the difference between the trajectorıes of Dk,k’’(τ) and Dk’,k’’(τ) in a
given time epoch. In this case, the ED plots for Dk,k’’(τ), Dk’,k’’(τ) and Dk,k’(τ) coincide with
each other about the zero. Therefore, the modes of the ED plots of Dk,k’’(τ) and Dk’,k’’(τ) are
far from or close to each other if that of Dk,k’(τ) is far from or close to zero, respectively. See,
the trajectories of various spatial differences in Figs. 7 (a) and (b), 11 (a)-(c), and 21 (a)-(c)
and their ED plots in Figs. 8 (a) and (b), 12 (a)-(d), and 20 (a), respectively.
- 15 -
APPENDIX
Suppose that the number of the samples X(i) in a finite set is N. Let us consider Eq. (1) in
the main text with P(X(i))=Q(X(i))/N where P(X(i))=P(i) and Q(X(i))=Q(i) are the
normalized and un normalized empirical distributions of X(i), respectively and 1 ≤ i ≤ N;
U = -N∑N
i=1P(i)log(NP(i))
(A1)
(A5)
(A3)
(A6)
(A2)
(A4)
U/N = S - log(N)
U = NS - Nlog(N)
U/(Nlog(N)) = S/log(N) - 1
1 + U/(Nlog(N)) = S/log(N) = χ
i=1(P(i)log(P(i) + P(i)log(N))
U = -N∑N
or
or
since P is normalized, ∑N
i=1P(i)=1. Thus,
or
or
where χ is the same as the index in Eq. (2) in the main text.
ACKNOWLEDGEMENT
The author is thankful to the University of Freiburg for their kindness in giving permission to
investigate their database and use all of the related material.
- 16 -
REFERENCE LIST
[1] F. H. Lopes da Silva, EEG analysis: theory and practice; in: E. Niedermeyer, F. H. Lopes
da Silva, eds. Electroencephalography: basic principles, clinical applications and related
fields, Baltimore: Williams and Wilkins, 3rd ed. 1993.
[2] F.H. Lopes da Silva, S. van Leeuwen, A. Réemond, Handbook of Electroencephalography
and Clinical Neurophysiology, Vol. II: Clinical Application of Computer Analysis of EEG
and Other Neurophysiological Signals, Elsevier, Amsterdam, 1986.
[3] F. H. Lopes Silva, A. Hoeks, T. H. M. T. van Lierop, C. F. Schrijer and W. S. van
Leeuwen, Confidence intervals of spectra and coherence functions—Their relevance for
quantifying thalamo-cortical relationships; in Quantitative Analysis of the EEG M. Matejcek
and G. K. Schenk, eds. AEG Telefunken, Constanz, 1975.
[4] P. Kellawou and I. Peterson, Automation of Clinical EEG, eds. New York: Raven, 1973.
[5] S. Faula, G. Boylanb, S. Connollyc, L. Marnanea, G. Lightbodya, An evaluation of
automated neonatal seizure detection methods, Clinical Neurophysiology 116, 1533, 2005.
[6] J. D. Bronzino, M. L. Kelly, C. T. Cordova, N. H. Oley, and P. J. M., Utilization of
Amplitude Histographs to Quantify the EEG Effects of Systemic Administration of Morphine
in the Chronically Implanted Rat, IEEE BME-28, 673, 1981.
[7] J. S. Barlow, Computerized Clinical Electroencephalography in Perspective, IEEE BME-
26, 377, 1979.
[8] A. S. Gevins, C. L. Yeager, S. L. Diamond, J.-P. Spire, G. M. Zeitlin, and A. H. Gevins,
Automated analysis of the electrical activity of the human brain (EEG): A progress report,
Proc. IEEE, vol. 63, 1382, 1975.
[9] Y.U. Khan and J. Gotman, Wavelet based automatic seizure detection in intracerebral
Electroencephalogram, Clinical Neurophysiology 114, 898, 2003.
[10] C. D. Binnie, B. G. Batchelor, P. A. Bowring, C. E. Darby, L. Herbert, D. S. L. Lloyd, D.
M. Smith, G. F. Smith, and M. Smith, Computer-assisted interpretation of clinical EEGs,
Electroencephalogr. Clinical Neurophysiology, 44, 575, 1978.
[11] H. W. Shipton, EEG analysis: A history and a prospective, Annu. Rev. Biophys. Bioeng.,
4, 1, 1975.
[12] N. V. Thakor and S. Tong, Advances in quantitative electroencephalogram analysis
methods. Annu. Rev. Biomed. Eng. 6, 453, 2004.
[13] I. Osorio, M-G. Frei, S. B. Wilkinson. Real-time automated detection and quantitative
analysis of seizures and short-term prediction of clinical onset, Epilepsia, 39, 615, 1998.
[14] T. Kreuz, F. Mormann, R. G. Andrzejak, A. Kraskov, K. Lehnertz, P. Grassberger,
Measuring synchronization in coupled model systems: A comparison of different approaches,
Physica D 225, 29, 2007.
[15] F. H. Lopes da Silva, W. Blanes, S. Kalitzin, J. Parra, P. Suffczynski, D. Velis,
Dynamical diseases of brain systems: different routes of epileptic seizures. IEEE BME-50,
540, 2003.
[16] R. G. Andrzejak, D. Chicharro, C. E. Elger, F. Mormann, Seizure prediction: Any better
than chance?, Clinical Neurophysiology 120, 1465, 2009.
[17] M. Le Van Quyen, J. Martinerie, C. Adam, F. Varela, Nonlinear analyses of interictal
EEG map the interdependencies in human focal epilepsy. Physica D, 127, 250, 1999.
[18] J. Arnhold, P. Grassberger, K. Lehnertz, C. Elger, A robust method for detecting
interdependencies: application to intracranially recorded EEG. Physica D, 134, 419, 1999.
[19] C. Stam, J. Pijn, P. Suffczynski, F. H. Lopes da Silva, Dynamics of the alpha rhythm:
evidence for non-linearity? Clinical Neurophysiology 110, 1801, 1999.
- 17 -
[20] F. Mormann, T. Kreuz, C. Rieke, R. G. Andrzejak, A. Kraskov, P. David, C. E. Elger,
Klaus Lehnertz, On the predictability of epileptic seizures, Clinical Neurophysiology 116,
569, 2005.
[21] F. Mormann, R. G. Andrzejak, C. E. Elger and K. Lehnertz, Seizure prediction: the long
and winding road, Brain, 130, 314, 2007.
[22] M. E. Weinand, L. P. Carter, W. F. El-Saadany, P. J. Sioutos, D. M. Labiner, K. J.
Oommen, Cerebral blood flow and temporal lobe epileptogenicity, J. Neurosurg, 86, 226,
1997.
[23] C. Baumgartner, W. Serles, F. Leutmezer, E. Pataraia, S. Aull, T. Czech, U. Pietrzyk, A.
Relic, I. Podreka, Preictal SPECT in temporal lobe epilepsy: regional cerebral blood flow is
increased prior to electroencephalography-seizure onset, J. Nucl. Med. 39, 978, 1998.
[24] P. D. Adelson, E. Nemoto, M. Scheuer, M. Painter, J. Morgan, H. Yonas, Noninvasive
continuous monitoring of cerebral oxygenation periictally using nearinfrared spectroscopy: a
preliminary report. Epilepsia 40, 1484, 1999.
[25] P. Federico P, D. F. Abbott, R. S. Briellmann, A. S. Harvey, G. D. Jackson, Functional
MRI of the pre-ictal state, Brain 128, 1811, 2005.
[26] R. Delamont, P. Julu, G. Jamal, Changes in a measure of cardiac vagal activity before
and after epileptic seizures, Epilepsy Res., 35, 87, 1999.
[27] V. Novak, A. L. Reeves, P. Novak, P. A. Low, F. W. Sharbrough, Time-frequency
mapping of R-R interval during complex partial seizures of temporal lobe origin. J. Auton
Nerv. Syst. 77, 195, 1999.
[28] D. H. Kerem and A. B. Geva, Forecasting epilepsy from the heart rate signal, Med. Biol.
Eng. Comput. 43, 230, 2005.
[29] E. Pereda, R. Q. Quiroga, J. Bhattacharya, Nonlinear multivariate analysis of
neurophysiological signals, Progress in Neurobiology, 77, 1, 2005.
[30] M.J. van der Heydena, D.N. Velis, B.P.T. Hoekstra, J.P.M. Pijn, W. van Emde Boas, C.
W. M. van Veelen, P.C. van Rijen, F.H. Lopes da Silva, J. De Goede, Non-linear analysis of
intracranial human EEG in temporal lobe epilepsy, Clinical Neurophysiology 110, 1726,
1999.
[31] T. Schreiber, Interdisciplinary application of nonlinear time series methods, Physics
Reports 308, 1, 1999.
[32] Freiburg University, http://epilepsy.uni-freiburg.de/prediction-contest/data-download
[33] B. Bein, Entropy, Best Practice & Research Clinical Anaesthesiology 20, 101, 2006.
[34] P. Viola, N. N. Schraudolph, T. J. Sejnowski, Empirical Entropy Manipulation for Real-
World Problems, in Advances in Neural Information Processing Systems 8 (NIPS*96), D.
Touretzky, M. Mozer and M. Wasselmo (eds.), MIT Press, 1996; pp.851-857.
[35] A. Isaksson, A. Wennberg, L. H. Zetterberg, Computer Analysis of EEG Signals with
Parametric Models, PROCEEDINGS OF THE IEEE, 69, 451, 1981.
[36] A. Bezerianos, S. Tong S, N. Thakor, Time-dependent entropy estimation of EEG
rhythm changes following brain ischemia, Ann. Biomed. Eng, 31, 221, 2003.
[37] C. E. Shannon, W. Weaver, The Mathematical Theory of Communication, University of
Illinois Press, Urbana, 1948.
[38] C. E. Shannon, Bell Systems Technol. J. 27, 379, 1948.
[39] S. M. Zoldi, A. Krystal, H. S. Greenside, Stationarity and Redundancy of Multichannel
EEG Data Recorded During Generalized Tonic-Clonic Seizures, Brain Topography, 12, 187,
2000.
[40] M. Breakspear and J.R. Terry, Detection and description of non-linear interdependence
in normal multichannel human EEG data, Clinical Neurophysiology 113, 735, 2002.
[41] J. S. Richman and J. R. Moorman, Physiological time-series analysis using approximate
entropy and sample entropy, Am. J. Physiol. Heart. Circ. Physiol. 278, 2039, 2000.
- 18 -
[42] O. Vasicek, A Test for Normality Based on Sample Entropy, Journal of the Royal
Statistical Society B, 38, 54, 1976.
[43] E. J. Dudewicz and E. C. van der Meulen, Entropy-Based Tests of Uniformity, Journal of
the American Statistical Association, 76, 967, 1981.
[44] S. Park, A goodness-of-fit test for normality based on the sample entropy of order
statistics, Statistics & Probability Letters, 44, 359, 1999.
[45] R. A.H. Lorentz, On the entropy of a function, Journal of Approximation Theory, 158,
145, 2009.
[46] J. Timmer, from Freiburg University (private communication).
[47] R. Aschenbrenner-Scheibe, T. Maiwald T, M. Winterhalder, H. U. Voss, J. Timmer, A.
Schulze-Bonhage, How well can epileptic seizures be predicted? An evaluation of a
nonlinear method, Brain, 126, 2616, 2003.
[48] M. Winterhalder, B. Schelter, T. Maiwald, A. Brandt, A. Schad, A. Schulze-Bonhage, J.
Timmer, Spatio-temporal patient-individual assessment of synchronization changes for
epileptic seizure prediction. Clinical Neurophysiology 117, 2399, 2006.
[49] M. Winterhalder, T. Maiwald T, H. U. Voss, R. Aschenbrenner-Scheibe, J. Timmer, A.
Schulze-Bonhage, The seizure prediction characteristic: a general framework to assess and
compare seizure prediction methods, Epilepsy Behav., 4(3), 318, 2003.
[50] T. Maiwald, M. Winterhalder, R. Aschenbrenner-Scheibe R, H. U. Voss, A. Schulze-
Bonhage, J. Timmer, Comparison of three nonlinear seizure prediction methods by means of
the seizure prediction characteristic, Physica D, 194, 2004.
[51] B. Schelter, M. Winterhalder, T. Maiwald, A. Brandt, A. Schad, J. Timmer, A. Schulze-
Bonhage.: Do false predictions of seizures depend on the state of vigilance? A report from
two seizure prediction methods and proposed remedies, Epilepsia, 47, 2058, 2006.
[52] B. Schelter, M. Winterhalder, T. Maiwald, A. Brandt, A. Schad, A. Schulze-Bonhage,
and J. Timmer Testing statistical significance of multivariate time series analysis techniques
for epileptic seizure prediction, Chaos 16, 013108, 2006.
[53] J. Bruhn, L. E. Lehmann, H. Ropcke, T. W. Bouillon, A. Hoeft, Shannon entropy applied
to the measurement of the electroencephalographic effects of desflurane, Anesthesiology, 95,
30, 2001.
[54] I. J. Rampil, A primer for EEG signal processing in anesthesia, Anesthesiology, 89, 980,
1998.
[55] W. D. Smith, R. C. Dutton, N. T. Smith, Measuring the performance of anesthetic depth
indicators, Anesthesiology 84, 38, 1996.
[56] D. T.J. Liley, K. Leslie, N. C. Sinclair, M. Feckie, Dissociating the effects of nitrous
oxide on brain electrical activity using fixed order time series modeling, Computers in
Biology and Medicine, 38, 1121, 2008.
[58] M. Le Van Quyen, J. Soss, V. Navarro, R. Robertson, M. Chavez, M. Baulac, J.
Martinerie, Preictal state identification by synchronization changes in long-term intracranial
EEG recordings, Clinical Neurophysiology 116, 559, 2005.
[59] S. Whiting, T. Ning, and J. D. Bronzino, Data length effects on the coherence estimate of
EEG, Bioengineering Conference, 1989., Proceedings of the 1989 Fifteenth Annual
Northeast, page 93, 1989.
[60] T. D. de Wit, When do finite sample effects significantly affect entropy
estimates? Eur. Phys. J. B 11, 513, 1999.
[61] M. A. Jime´nez-Montano, W. Ebeling, T. Pohl, Paul E. Rapp, Entropy and complexity of
finite sequences as fluctuating quantities, BioSystems 64, 23, 2002.
- 19 -
TABLES
CP,GTC
Pat11 10,f
Pat14 41,f
Frontal
G_A4, IH4, IH3; G_D2, IHA1, IH1
Pat01
SP,CP
15,f
Pat03 14,m SP,CP
FE1, FE2, FD1; G_F8, G_G8, G_H8
SP,CP,GTC G_A3, G_A4, IHB4; G_D1, IHB1, LL1
Pat05 16,f
FRA2, FRA3, FRC1; G_A8, FLA4, FLC6
Pat08 32,f
SP,CP
Pat18 25,f
G_A7, G_C5, G_E5; IHC4, IHB4, G_F3
SP,CP
SP,CP,GTC G_B5, G_D8, G_G7; G_C1, G_A1, G_H4
Pat19 28,f
Fronto/Temporal
FRA1, FRB1, TRB2; TRA1, TRC1, TRC6
Tempo/Parietal
Pat20 33,m SP,CP,GTC G_D3, G_B4, G_A2; H_5, FLB4, TLC1
Parietal
SP,CP,GTC G_G3, G_G4, G_E3; G_C8, G_D8, G_F7
Temporal
SP,CP,GTC HR_2, HR_5, TBB1; HR_9, G_D6, G_A1
Pat04
26,f
Pat07
SP,CP,GTC TLA1, TLB2, TLC2; TLA5, TLB5, TLC6
42,f
Pat10 47,m SP,CP,GTC TLA1, TLB1, TLB2; TRB2, TRC2, TRC5
Pat12 42,f
SP,CP,GTC TBA4, TBB6, HR_7; TLB2, TLB3, TLC2
Pat15 31,m SP,CP,GTC TBA1, TLR4, TLA4; HR_8, FRA6, FRB3
Pat16 50,f
SP,CP,GTC HL_2, HL_3, TBB1; G_H1, HL_9, TBC2
Pat17 28,m SP,CP,GTC TBA1, TBA2, TLA1; G_B5, G_B8, G_C5
TBB1, G_E6, TBA3; G_A2, G_B2, G_C2
Pat21 13,m SP,CP
Temporo/Occipital
Pat06 31,f
TLC1, TLC2, OBB1; TRC1, GD8, OBB4
CP,GTC
Pat09 44,m CP,GTC
TBA3, TBB2, TBC4; G_A1, G_A8, TBC1
Pat13 22,f
SP,CP,GTC TBC1, TBC2, TBB1; POA1, POB1, TRA1
Table I Brief information about the patients with focal epilepsies [32] grouped with regard
to the origins of the epilepsies. The patients are designated on the left column (PatNM) and
the numbers on the second column show the ages where f or m stands for female or male,
respectively. The seizure types are shown on the third column which are simple partial (SP),
complex partial (CP) or generalized tonic-clonic (GTC). The first three of the electrodes given
on the fourth column are in-focus and the others (after the semicolons) are out-focus.
- 20 -
FIGURES
Figure 1
The available electrode types and contact positions of the declared patients
where the red and blue dots indicate in-focus or out-focus electrodes, respectively [32].
- 21 -
1.0
0.9
k
χ
0.8
1.0
0.9
k
χ
0.8
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
1M
100k
10k
1k
100
z
Pat03
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
1M
100k
10k
1k
100
z
0.7
10
0.7
3
2
1
0
8
7
6
5
4
τ (hour)
10 12 14 16 18 20 22 24
τ (hour)
(b)
(a)
Figure 2
The time profiles of the entropy indices for the available [32] ictal (a) and
inter-ictal (b) data from the Pat03 (see, Table I) where the arrows in (a) indicate the seizure
terms. The black line at the bottom is for z(τ) in the plots.
0
2
4
6
8
1.0
0.9
k
χ
0.8
1.0
0.9
k
χ
0.8
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
1M
100k
10k
1k
100
z
Pat03
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
1M
100k
10k
1k
100
z
0.7
10
0.7
2
0
1
τ (hour)
21
τ (hour)
(b)
(a)
Figure 3
The plots (a) and (b) are same as Figs. 2 (a) and (b), respectively where the
time domains are different.
20
22
10
10
1.0
0.9
k
χ
0.8
0.7
2
Figure 4
1.0
k
χ
0.9
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
1M
100k
10k
1k
100
z
1.0
0.9
k
χ
0.8
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
1M
100k
10k
1k
100
z
10
4
3
τ (hour)
τ (hour)
(b)
(a)
The plots are same as Figs. 2 (a) with different time domains.
5
10
6
0.7
4
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.08
k
'
χ
0.00
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.8
1.4
1.5
1.6
1.7
τ (hour)
(a)
1.8
1.9
2.0
-0.08
1.4
1.5
1.6
1.8
1.9
2.0
1.7
τ (hour)
(b)
- 22 -
0.14
0.00
k
'
'
χ
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
-0.14
1.4
1.8
1.9
1.5
2.0
1.7
1.6
τ (hour)
(c)
Figure 5
The time series of χk(τ) is depicted in (a) which is same as with 3 (a) for a
shorter time domain. The successive time differences χ’k(τ) and χ’’k(τ) are displayed in (b)
and (c), respectively with the same time domain as in (a), here.
1k
100
10
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
1
0.8
Pat03
ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.9
χ
1.0
1k
100
10
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
1
0.8
Pat03
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.9
χ
1.0
(a)
(b)
Figure 6
The distributions of the entropy index values for all of the data from the patient
in ictal interval (a) and in inter-ictal interval (b).
0.1
'
k
,
k
D
0.0
Pat03
ictal
D1,2
D1,3
D2,3
0.1
0.0
'
k
,
k
D
Pat03
ictal
D4,5
D4,6
D5,6
-0.1
1.0
1.2
1.4
1.6
τ (hour)
1.8
2.0
(a)
-0.1
1.0
1.2
1.4
1.6
τ (hour)
1.8
2.0
(b)
0.1
0.0
'
k
,
k
D
Pat03
ictal
D4,5
D4,6
D5,6
-0.1
0
1
2
3
5
4
τ (hour)
6
7
8
(c)
- 23 -
Figure 7
The synchronous differences of the entropy index values of the in-focus
electrodes (a) and out-focus electrodes (b) and (c) of the patient in ictal interval.
Pat03
ictal
D1,2
D1,3
D2,3
D4,5
D4,6
D5,6
0.00
0.08
1k
100
10
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
1
-0.1
Pat03
inter-ictal
D1,2
D1,3
D2,3
D4,5
D4,6
D5,6
0.0
0.1
Dk,k'
Dk,k'
(a)
(b)
Figure 8
The distribution of the values of the synchronous spatial differences of the in-
focus electrodes of the Pat03 in ictal interval (a) and in inter-ictal interval (b).
1k
100
10
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
1
-0.08
0
1
2
1.0
k
χ
0.9
0.8
0.7
0.6
1.0
k
χ
0.9
0.8
0.7
0.6
4
5
6
τ (hour)
(d)
(c)
Figure 9
Entropy index profiles from the Pat14 in ictal interval and in inter-ictal interval
are shown in (a)-(c) and (d), respectively with different time domains.
21.3
21.4
τ (hour)
21.5
1k
100
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
10
1
Pat14
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.6
0.7
0.8
χ
0.9
1.0
Pat14
ictal
k=1
k=2
k=3
k=4
k=5
k=6
1k
100
10
1
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
- 24 -
0.6
0.7
0.8
χ
0.9
1.0
10k
1k
100
Pat14
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10
z
1
1.0
k
χ
0.9
0.8
0.7
0.6
1
τ (hour)
(b)
6
7
5
4
3
τ (hour)
(a)
1.0
k
χ
0.5
0.0
21.2
10k
1k
100
Pat14
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10
z
1
7
10k
1k
100
Pat14
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10
z
1
2
Pat14
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
(a)
(b)
Figure 10
The ED of the index values of the EEG from the Pat14 in ictal interval (a) and
in inter-ictal interval where the tails are noisy due to several irregular behaviors in the data.
0.18
0.09
'
k
,
k
D
0.00
Pat14
ictal
D1,5
D2,3
D2,4
D2,5
0.1
'
k
,
k
D
0.0
-0.1
Pat14
ictal
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
-0.09
3.0
3.2
3.4
3.6
τ (hour)
3.8
4.0
3.0
3.2
3.4
3.6
τ (hour)
3.8
4.0
(b)
(a)
0.2
'
k
,
k
D
0.0
-0.2
5
6
τ (hour)
7
Pat14
ictal
D1,2
D1,3
D1,4
D1,5
D1,6
D2,3
D2,4
D2,5
D2,6
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
(c)
Figure 11
The time profiles of several local and non-local electrode couplings about the
same (a) and (b) or different seizure terms (c).
100
10
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
1
-0.1
Pat14
ictal
D1,2
D1,3
D1,5
D2,3
D2,5
0.0
0.1
Dk,k'
0.2
100
10
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
1
-0.2
Pat14
ictal
D4,5
D4,6
D5,6
-0.1
Dk,k'
0.0
0.1
(a)
(b)
- 25 -
100
10
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
1
-0.2
Pat14
ictal
D1,2
D1,3
D2,3
D4,5
D4,6
D5,6
1k
100
10
1
-0.2
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
-0.1
0.0
Dk,k'
0.1
Pat14
inter-ictal
D1,2
D1,3
D2,3
D4,5
D4,6
D5,6
-0.1
0.0
Dk,k'
0.1
0.2
(d)
(c)
Figure 12
Various cross-correlations between the in-focus and out-focus electrodes of
the patient in ictal interval (a)-(c) and in inter-ictal interval (b).
1.0
k
χ
0.9
0.8
0.7
1.0
k
χ
0.9
0
1
2
3
4
5
9 10 11 12 13
8
6
7
τ (hour)
(a)
1.0
k
χ
0.9
0.8
0.7
0
1
2
τ (hour)
(b)
100k
10k
1k
Pat20
ictal
k=1
k=2
k=3
k=4
k=5
k=6
100
z
10
3
100k
10k
1k
Pat20
ictal
k=1
k=2
k=3
k=4
k=5
k=6
100
z
10
Pat20
ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.8
1.8
2.0
2.2
2.4
τ (hour)
(c)
Figure 13
The time profiles of the indices of the Pat20 in ictal interval with different time
domains (a)-(c).
1k
100
10
1
0.8
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
Pat20
ictal
k=1
k=2
k=3
k=4
k=5
k=6
1k
100
10
1
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
Pat20
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.9
χ
1.0
0.8
0.9
χ
1.0
(b)
(a)
Figure 14
The ED plots for the index values from the Pat20 in ictal interval (a) and in
inter-ictal interval (b).
- 26 -
1.0
k
χ
0.9
0.8
0.7
1.0
k
χ
0.9
1.0
k
χ
0.9
Pat11
ictal
k=1
k=2
k=3
k=4
k=5
k=6
100k
10k
1k
z
100
Pat11
ictal
k=1
k=2
k=3
k=4
k=5
k=6
100k
10k
1k
z
100
0
1
2
3
(a)
4
τ (hour)
5
6
7
8
0.8
2
3
5
6
4
τ (hour)
(b)
Pat11
ictal
k=1
k=2
k=3
k=4
k=5
k=6
100k
10k
1k
z
100
0.8
6
8
7
τ (hour)
(c)
Figure 15
The time profiles of the ictal entropy indices of the Pat11 in (a)-(c) with
various time domains.
1k
100
10
1
0.8
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
Pat11
ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.9
χ
1.0
1k
100
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
10
1
0.8
Pat11
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.9
χ
1.0
(b)
(a)
Figure 16
The ED plots of the index values of the Pat11 in ictal interval (a) and in inter-
ictal-interval (b).
- 27 -
1k
100
10
1
-0.05
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
Pat11
ictal
D1,2
D1,3
D1,4
D1,5
D1,6
D2,3
D2,4
D2,5
D2,6
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
1k
100
10
1
-0.05
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
Pat11
inter-ictal
D1,2
D1,3
D1,4
D1,5
D1,6
D2,3
D2,4
D2,5
D2,6
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
0.00
Dk,k'
0.05
0.00
Dk,k'
0.05
(b)
(a)
Figure 17
The local and non-local cross-correlations between the electrodes of the Pat11
in ictal interval (a) and in inter-ictal interval (b).
Pat15
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10M
1M
100k
10k
1k
100
z
10
0
1
2
3
5
4
τ (hour)
6
7
8
9
1
10
(a)
1.0
k
0.9
χ
0.8
0.7
0.6
0.5
0
1.0
k
0.9
χ
0.8
0.7
0.6
1
2
τ (hour)
3
(b)
Pat15
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10M
1M
100k
10k
1k
100
z
10
1
4
Pat15
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10M
1M
100k
10k
1k
100
z
10
0.5
7
1
10
Pat15
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10M
1M
100k
10k
1k
100
z
10
1
1.0
k
0.9
χ
0.8
0.7
0.6
0.5
1.0
k
0.9
χ
0.8
0.7
0.6
0.5
4
5
6
7
τ (hour)
(d)
(c)
Figure 18
The time profiles of the index values from the Pat15 in ictal interval with
different time domains.
τ (hour)
8
9
1k
100
10
1
0.6
)
2
0
0
.
0
=
χ
δ
(
χ
f
o
s
t
n
u
o
c
Pat15
ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.7
0.8
χ
0.9
1.0
(a)
1k
100
10
1
0.7
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
- 28 -
Pat15
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.8
0.9
χ
1.0
(b)
Figure 19
The ED plots of the index values from the Pat15 in ictal interval (a) and in
inter-ictal interval.
1k
100
10
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
1
-0.15
-0.10
-0.05
0.00
Dk,k'
0.05
0.10
0.15
Pat15
ictal
D1,2
D1,3
D1,4
D1,5
D1,6
D2,3
D2,4
D2,5
D2,6
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
1k
100
)
1
0
0
.
0
=
D
δ
(
s
t
n
u
o
c
10
1
-0.15
-0.10
-0.05
0.00
Dk,k'
0.05
0.10
0.15
Pat15
inter-ictal
D1,2
D1,3
D1,4
D1,5
D1,6
D2,3
D2,4
D2,5
D2,6
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
(b)
(a)
Figure 20
The cross-correlations between the electrodes of the Pat15 in ictal interval (a)
and in inter-ictal interval (b).
0.2
0.1
'
k
,
k
D
0.0
-0.1
-0.2
5.4
5.5
5.6
5.7
τ (hour)
5.8
5.9
(b)
Pat15
ictal
D1,2
D1,3
D1,4
D1,5
D1,6
D2,3
D2,4
D2,5
D2,6
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
0.1
0.0
'
k
,
k
D
-0.1
-0.2
4
0.2
0.1
'
k
,
k
D
0.0
-0.1
-0.2
8.4
Pat15
ictal
D4,5
D4,6
D5,6
Pat15
ictal
D1,2
D1,3
D1,4
D1,5
D1,6
D2,3
D2,4
D2,5
D2,6
D3,4
D3,5
D3,6
D4,5
D4,6
D5,6
5
6
τ (hour)
(a)
7
8.6
8.8
τ (hour)
9.0
(c)
Figure 21
The cross-correlations between the out-focus electrodes (a) and all of the
electrodes ((b) and (c)) of the Pat15 in ictal interval within different time domains.
- 29 -
10k
100k Pat09
ictal
k=1
k=2
k=3
k=4
k=5
k=6
1k
0
1
2
3
4
5
τ (hour)
(a)
6
7
8
9
10
100
10
1
Pat09
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
1.0
k
0.9
χ
0.8
0.7
0.6
0.5
1.0
k
χ
0.9
0.8
8
9
τ (hour)
10
11
1k
100
10
1
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
Pat09
ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.8
0.9
χ
1.0
(c)
Figure 22
The time profiles of the index values from the Pat09 in ictal interval (a) and (b)
and in inter-ictal interval (c) with different time domains.
(b)
(a)
Figure 23
ED plots of the index values from the Pat09 in ictal interval (a) and in inter-
ictal interval (b).
1.0
k
0.9
χ
0.8
0.7
0.6
0.5
8
100k
1M Pat09
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10k
1k
100
10
10
9
τ (hour)
(b)
1k
100
10
1
0.7
)
1
0
0
.
0
=
χ
δ
(
s
t
n
u
o
c
Pat09
inter-ictal
k=1
k=2
k=3
k=4
k=5
k=6
0.8
0.9
χ
1.0
- 30 -
100k
)
2
0
0
.
0
=
'
χ
δ
(
'
χ
f
o
s
t
n
u
o
c
10k
1k
100
10
1
k=1
k=2
k=3
k=4
k=5
k=6
-1.0
-0.5
χ'
0.0
0.5
1.0
Figure 24
The distributions of the values χ’k(τ) which are obtained using all of the data
from 20 patients in ictal interval or in inter-ictal interval.
1.0
0.9
k
χ
0.8
0.7
0.6
0.5
0
0.2
k
'
χ
0.0
1
2
τ (hour)
(a)
100k Pat20
ictal
10k
k=1
k=2
k=3
k=4
k=5
k=6
100
1k
10
1
z
100m
3
0.8
k
χ
0.7
0.6
6.0
6.2
Pat20
ictal
k=1
k=2
k=3
k=4
k=5
k=6
100
1k Pat20
ictal
k=1
k=2
k=3
k=4
k=5
k=6
10
1
z
6.8
100m
7.0
6.4
6.6
τ (hour)
(b)
-0.2
1.0
1.2
2.0
1.8
1.4
1.6
τ (hour)
(c)
Figure 25
Various applications of the method with a window time length of 100 seconds
where (a) and (b) show the indices and (c) is for the temporal differences of the indices.
- 31 -
|
1401.0165 | 1 | 1401 | 2013-12-31T16:00:53 | Orthogonal invariant sets of the diffusion tensor and the development of a curvilinear set suitable for low-anisotropy tissues | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.QM"
] | We develop a curvilinear invariant set of the diffusion tensor which may be applied to Diffusion Tensor Imaging measurements on tissues and porous media. This new set is an alternative to the more common invariants such as fractional anisotropy and the diffusion mode. The alternative invariant set possesses a different structure to the other known invariant sets; the second and third members of the curvilinear set measure the degree of orthotropy and oblateness/prolateness, respectively. The proposed advantage of these invariants is that they may work well in situations of low diffusion anisotropy and isotropy, as is often observed in tissues such as cartilage.
We also explore the other orthogonal invariant sets in terms of their geometry in relation to eigenvalue space; a cylindrical set, a spherical set (including fractional anisotropy and the mode), and a log-Euclidean set. These three sets have a common structure. The first invariant measures the magnitude of the diffusion, the second and third invariants capture aspects of the anisotropy; the magnitude of the anisotropy and the shape of the diffusion ellipsoid (the manner in which the anisotropy is realised). We also show a simple method to prove the orthogonality of the invariants within a set. | physics.bio-ph | physics |
Orthogonal Invariant Sets of the Diffusion Tensor and the
Development of a Curvilinear Set Suitable for
Low-Anisotropy Tissues ∗
Robin A. Damion1,2
, Aleksandra Radjenovic3,†
, Eileen Ingham2
,
Zhongmin Jin2,4
, Michael E. Ries1,††
1 School of Physics and Astronomy,
University of Leeds, Leeds, West Yorkshire, UK.
2 Institute of Medical and Biological Engineering,
School of Mechanical Engineering, University of Leeds, Leeds, West Yorkshire, UK.
3 National Institute for Health Research, Leeds Musculoskeletal Biomedical Research Unit,
Chapel Allerton Hospital, University of Leeds, Leeds, West Yorkshire, UK.
4 School of Mechanical Engineering,
Xi'an Jiaotong University, Xi'an, Shaanxi, People's Republic of China.
† E-mail: [email protected]
†† E-mail: [email protected]
Abstract
We develop a curvilinear invariant set of the diffusion tensor which may be applied to Diffusion
Tensor Imaging measurements on tissues and porous media. This new set is an alternative to the
more common invariants such as fractional anisotropy and the diffusion mode. The alternative
invariant set possesses a different structure to the other known invariant sets; the second and
third members of the curvilinear set measure the degree of orthotropy and oblateness/prolateness,
respectively. The proposed advantage of these invariants is that they may work well in situations
of low diffusion anisotropy and isotropy, as is often observed in tissues such as cartilage.
We also explore the other orthogonal invariant sets in terms of their geometry in relation to
eigenvalue space; a cylindrical set, a spherical set (including fractional anisotropy and the mode),
and a log-Euclidean set. These three sets have a common structure. The first invariant measures
the magnitude of the diffusion, the second and third invariants capture aspects of the anisotropy;
the magnitude of the anisotropy and the shape of the diffusion ellipsoid (the manner in which the
anisotropy is realised). We also show a simple method to prove the orthogonality of the invariants
within a set.
∗Published in PLoS ONE, November 2013. This article is reproduced here because of formatting errors by PLoS ONE
in the post-production article. PLoS ONE 8(11): e78798. doi:10.1371/journal.pone.0078798
1
Invariant Sets of the Diffusion Tensor
Introduction
Fluids in partially-ordered biological tissues such as brain [1], muscle [2,3] and cartilage [4 -- 6] have been
shown to exhibit anisotropic translational diffusion when measured with techniques such as diffusion
tensor imaging (DTI) [7]. The fluids, which usually consist primarily of water, are inherently isotropic
liquids and therefore any measured anisotropy in their translational motion is a result of the influence
of a locally anisotropic environment on the motional statistics of the liquid molecules. The fact that
the anisotropy is caused by the local environment is useful because it provides a source of additional
contrast in magnetic resonance images and it is often possible to infer something about the local
microstructure of the tissue (or porous medium) from the anisotropy of the diffusion tensor and its
eigenvectors.
Normalised eigenvectors provide information about the local orientation of structures in the tissue,
while the anisotropy (a comparison of the eigenvalues) provides a measure of the relative magnitudes
of the eigenvectors. However, there is no single definition of anisotropy. Currently, the most common
measure of diffusion anisotropy is fractional anisotropy [8]
Af = vuut
3
2P3
i=1(cid:0)λi − ¯λ(cid:1)2
P3
i=1 λ2
i
,
(1)
Pi λ2
i which provides a scale-independent, dimensionless measure.
(displayed in its most familiar form) where λi (i = 1, 2, 3) are the three eigenvalues of the diffusion
tensor, and ¯λ = (λ1 + λ2 + λ3)/3 is the mean of the eigenvalues. Fractional anisotropy quantifies
anisotropy by the spread (standard deviation) of the eigenvalues and is normalised by the quantity
Importantly, and unlike some
previous attempts to define anisotropy [8], Af is invariant. That is, the quantity is independent of the
choice of coordinate system; in particular, with regard to rotations, but also more general coordinate
transformations. This coordinate independence is a consequence of the application of tensor calculus
which can be used to express Af , as was done by Basser and Pierpaoli [8]. Invariant quantities derived
from tensor calculus are typically symmetric functions of the eigenvalues, as can be seen in Eq. (1).
Some other invariant measures that have been proposed in the literature are relative anisotropy [7,
8], geodesic anisotropy [9, 10], and volume ratio [7]. The first two of these are similar to fractional
anisotropy in the sense that they are based on the spread of the eigenvalues. Volume ratio is a little
different in structure and is based on the volume of the diffusion ellipsoid which is proportional to the
determinant of the diffusion tensor. However, like fractional anisotropy, these quantities do not give
us the full picture of the anisotropy. Except for volume ratio, these quantities calculate, in different
ways, measures of the magnitude of the anisotropy but give us little idea of the shape of the diffusion
ellipsoid. Volume ratio gives a single mixed measure of anisotropy magnitude and ellipsoid shape [11].
The shape of the diffusion ellipsoid can be generally classified by four archetypal cases or states;
isotropic, orthotropic, prolate, and oblate, although only three are independent. For these, sets of
eigenvalues can take the form
{λ1, λ2, λ3} =
(cid:8)¯λ, ¯λ, ¯λ(cid:9) ,
(cid:8)¯λ (1 + α) , ¯λ, ¯λ (1 − α)(cid:9) ,
(cid:8)¯λ (1 + 2α) , ¯λ (1 − α) , ¯λ (1 − α)(cid:9) ,
(cid:8)¯λ (1 + α/2) , ¯λ (1 + α/2) , ¯λ (1 − α)(cid:9) , Oblate
Isotropic
Orthotropic
Prolate
(2)
PLoS ONE
2
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
where α ∈ (0, 1). Note that these states are parameterized by α strictly on the open interval. That is,
α = 1 should be regarded as an asymptotic limit because at this point one or more eigenvalues would
be zero and the determinant of the diffusion tensor would vanish, rendering the tensor noninvertible.
At the other end of the interval α = 0 and the states become a single degenerate state; the isotropic
case.
Fractional anisotropy does not easily distinguish between the states indicated above. For prolate
ellipsoids, Af ∈ (0, 1), for oblate ellipsoids Af ∈ (0, 1/√2), and for the orthotropic case Af ∈ (0,p3/5).
Thus, only when Af ∈ [p3/5, 1) can we identify the ellipsoid shape: prolate. In principle, the isotropic
case is also identifiable as Af = 0 but experimental noise [12] inevitably places a lower limit on the
measured value of fractional anisotropy, meaning that Af = 0 is unachievable in practice.
There have been some attempts to define quantities which describe the shape of the ellipsoid.
For example, Peled et al. [13] and Westin et al. [14] defined three quantities to measure the linear
(prolate), planar (oblate) and spherical (isotropic) aspects of the diffusion tensor. These quantities
were defined as functions on magnitude-ordered eigenvalue sets and therefore the functions treated the
eigenvalues asymmetrically. The main difficulty with this approach is that averaging of magnitude-
ordered eigenvalues usually introduces a statistical bias in the distribution of the eigenvalues, leading
to inaccurate measurements of anisotropy [15, 16].
A solution to the problem of finding invariant measures of anisotropy which encode the shape of
the diffusion ellipsoid, and which do not depend on having to magnitude-order the eigenvalues, was
proposed by Ennis and Kindlmann [17]. These authors, basing their work on that of Criscione et
al. [18], showed that there exist orthogonal sets (triplets) of tensor invariants which can be applied
to the diffusion tensor to give useful measures of anisotropy. They found two sets of invariants which
are related to cylindrical and spherical coordinates in eigenvalue space. The latter set was shown to
contain fractional anisotropy as a measure of the magnitude of anisotropy, and an invariant known as
the mode; a measure of the shape of the diffusion ellipsoid.
In the next section of this paper we review the orthogonal cylindrical and spherical sets using a
more geometric approach than was used by Ennis and Kindlmann. These authors gave a proof of the
orthogonality of these measures in the same way as Criscione et al. but this proof is based on tensors
in the full six-dimensional space of the diffusion tensor and the connection with the coordinate system
of the eigenvalue space is obscured. Here, we give an alternative proof of the orthogonality of these
invariant measures which maintains a direct connection with the eigenvalue coordinates and therefore
presents a clearer picture of the geometry and the meaning of the invariants. In fact the geometrical,
coordinate-based approach we take here is similar to that used by Bahn [11] prior to the work of the
abovementioned authors. Bahn was able to show how the spherical coordinates of eigenvalue space
were invariant themselves or related to the invariants forming an orthogonal set. What he did not
show was how these invariants were related to invariants of the full diffusion tensor and also explicitly
how his skewness parameter (related to the mode) determined the shape of the ellipsoid, nor did he
fully explore the invariants relating to the cylindrical set.
We also look at two further orthogonal sets. One set is related to recent work using the log-
Euclidean metric for averaging and interpolation of the diffusion tensor [9, 10]. The final set, the
curvilinear set, has a different structure to the previous three and makes the orthotropic ellipsoid case
more apparent. The incentive behind the definition of the latter set was our discovery that the third
invariants (measuring the mode or its equivalent) of the previous three sets (spherical, cylindrical,
log-Euclidean) do not function well for diffusion exhibiting isotropy or a low degree of anisotropy.
PLoS ONE
3
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
The primary purpose of this paper is to introduce the curvilinear invariant set and relate it and
compare it mathematically to the other invariant sets. It is not our intention to carry out a numerical or
experimental validation of the curvilinear invariants here but to present their definition, mathematical
properties, and to propose them as potentially useful invariants which warrant further study.
Theory
Orthogonality and the Metric
To prove the orthogonality of invariants, we are going to treat them as if they were the coordinates
on some manifold M and use the metric tensor gij, where i, j ∈ {1, 2, . . . , n}, and n is the dimension
of the manifold. The metric is a symmetric tensor which is used to calculate the distance along a
specified path on a manifold. Here, we shall only be concerned with infinitesimal distances ds given
by (using the Einstein summation convention)
ds2 = gijdxidxj,
(3)
where dxi are infinitesimal changes in the coordinates xi ∈ M. This is merely an alternative way of
displaying the metric tensor. For our purposes, the significance of using the metric is that for real
orthogonal coordinates the metric is diagonal and hence there will be only terms of the form gii(dxi)2
and no terms of the form gijdxidxj with i 6= j.
Let µ : M → N be a map from manifold M to N (which, for simplicity, we will assume to
be of the same dimension as M) such that µ : xi 7→ X i ∈ N .
If N is the same manifold as M,
µ : M → M is merely a coordinate transformation on M. To calculate the new metric γij on N ,
j = ∂xi/∂X j . The new metric and the orthogonality can
we need the Jacobian matrix J such that J i
then be assessed by substituting dxi = J i
j dX j into Eq. (3) or by computing the new metric directly
by γ = JTgJ, where JT is the transpose of the Jacobian matrix. In index notation γij = gklJ k
i J l
j.
Although coordinates are usually designated with superscript indices, as we have employed in
the above two paragraphs, below for convenience we will use subscripted indices to avoid confusion
between indices and exponents. This should not lead to any ambiguity since, from here on, index
notation will not be used extensively.
Cartesian Coordinates
Since tensor invariants are scalar functions of eigenvalues only, we will consider the three-dimensional
eigenvalue space only. We begin with the eigenvalues as orthogonal coordinates in three-dimensional
Euclidean space E. Strictly, since the eigenvalues must be positive real numbers, they reside only in
the open subset that is one octant of the full Euclidean space such that λ1, λ2, λ3 > 0. The Euclidean
metric tensor is given by
and the infinitesimal distance element is given by
g =
1 0 0
0 1 0
0 0 1
,
ds2 = gijdλidλj = dλ2
1 + dλ2
2 + dλ2
3.
(4)
(5)
PLoS ONE
4
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
However, it will be convenient to use a slightly different coordinate system defined, for example,
by
x
y
z
=
1/√6
1/√6 −p2/3
−1/√2 1/√2
0
1/√3
1/√3
1/√3
λ1
λ2
λ3
,
(6)
where the matrix in this equation is an orthogonal matrix (a rotation matrix similar to that used
by Bahn [11]) and has been chosen because it preserves the orthogonality of the coordinates while
defining the z-direction to be proportional to the tensor trace. The component of the rotation about
the z-axis, however, is arbitrary. The matrix shown in Eq. (6) is the inverse Jacobian matrix, as we
have defined it above, and since in this case it is also an orthogonal matrix, J−1 = JT.
The convenience of this choice is that in the case of isotropy, when the eigenvalues are identical,
x = y = 0 and therefore isotropic tensors will be represented by points on the z-axis and the anisotropy
is dependent on the x and y coordinates only. Note that while the z-coordinate, being proportional
to the tensor trace, is an invariant, the x and y coordinates are not invariants and therefore we will
need either to find another coordinate system or appropriate functions of the coordinates. However,
this Cartesian system will serve as an important first step in establishing the geometry.
In defining our new Cartesian coordinates, we used a rotation matrix. Therefore, we know that
the new coordinates will also be orthogonal and the standard Euclidean metric will be preserved.
Calculating the new metric,
ds2 = dx2 + dy2 + dz2,
(7)
we verify the orthogonality of these coordinates by the absence of off-diagonal terms such as dxdy.
Cylindrical Coordinates and Invariants
Let us change to a cylindrical polar coordinate system {z, ρ, θ} such that the z-axis remains unchanged
and
x = ρ cos θ,
y = ρ sin θ,
(8)
(9)
(10)
and the metric becomes
ds2 = dz2 + dρ2 + ρ2dθ2.
The cylindrical invariant set [17] can now be defined in terms of these coordinates by
K1 =
K2 =
√3z
ρ
= tr(D),
= norm( D),
K3 = − cos 3θ = 3√6
det( D)
norm( D)3
,
where D is the diffusion tensor, D = D − 1
3 tr(D)I is the deviatoric tensor, I is the identity tensor,
tr is the trace, det is the determinant, and norm(A)2 = tr(A2) for an arbitrary tensor A. The trace,
norm and determinant are tensor invariants and thus Eq. (10) can be defined directly in terms of the
PLoS ONE
5
2013 Volume 8 Issue 11 e78798
eigenvalues of D. In terms of the eigenvalues, the basic tensor invariants used here and below are
Invariant Sets of the Diffusion Tensor
i=1 λi,
tr(D) = P3
norm(D) = qP3
norm( D) = qP3
det( D) = Q3
i=1 λ2
i ,
i=1(λi − ¯λ)2,
i=1(λi − ¯λ),
(11)
(of which only three are independent) and the relations in (10) between the cylindrical invariants and
the cylindrical coordinates can be verified using (6), (8), and (11).
The cylindrical invariants have the following interpretation. K1 ∈ (0,∞) is proportional to the
tensor trace and is a measure of the magnitude of the diffusion tensor. K2 and K3, being functions on
the plane perpendicular to the z-axis, quantify the anisotropy. Like fractional anisotropy, K2 ∈ (0,∞)
measures the magnitude of the anisotropy by the spread of the eigenvalues. However, unlike fractional
anisotropy, K2 is not normalised and therefore not scale-independent. In the cylindrical coordinate
system K2 has the direct interpretation of being the perpendicular distance between point (z, ρ, θ)
and the z-axis; i.e. the radius ρ. K3 ∈ [−1, 1] is known as the mode and quantifies the shape of
the diffusion ellipsoid. For the three ellipsoid shapes given by (2), K3 = 1, 0,−1 for the prolate,
orthotropic and oblate cases respectively. In the case of isotropic ellipsoids, a little care needs to be
taken because K3 is not defined and depends on how this point (the z-axis) is approached.
For example, let us parameterize the prolate and oblate cases with a single parameter β ∈ (− 1
2 , 1)
via
{λ1, λ2, λ3} = (cid:8)¯λ (1 + 2β) , ¯λ (1 − β) , ¯λ (1 − β)(cid:9) ,
(12)
where β ∈ (− 1
2 , 0) corresponds to oblate ellipsoids and β ∈ (0, 1) corresponds to the prolate case.
Calculating the mode as a function of β, K3 = β3/β3 = sgn(β), and it is clear that a discontinuous
jump occurs as we pass through the z-axis (β = 0). A similar discontinuous jump occurs along
any radial path through the z-axis except for those angles, θ, which correspond to the orthotropic
ellipsoids. In these cases, the mode is zero along the whole radial path.
The invariant set (10) can be regarded as a triplet of scalar functions on the eigenvalue space with
the metric (9) but it is also possible to regard these functions as coordinates on a related manifold K.
Because the function K3 : θ → − cos 3θ is a six-to-one mapping, the eigenvalue manifold is a six-fold
covering of the manifold associated with coordinates {Ki}. This is a consequence of the invariance
of the {Ki} which treats the eigenvalues symmetrically; there are six equivalent permutations of
(λ1, λ2, λ3). The metric on K is
ds2 =
1
3
dK 2
1 + dK 2
2 +
K 2
2
9(1 − K 2
3 )
dK 2
3 .
(13)
Again, the absence of off-diagonal terms dKidKj verifies the orthogonality of the coordinates and
hence the set {Ki} are orthogonal tensor invariants. We also note that despite the singularity in the
metric on the boundary K3 = ±1, the metric is completely flat. That is, the Ricci scalar curvature
vanishes everywhere.
PLoS ONE
6
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
Spherical Coordinates and Invariants
The invariant K2 is similar to fractional anisotropy but lacks a normalisation which makes it scale-
independent. Ennis and Kindlmann [17] showed that it is possible to modify the set to include
fractional anisotropy and that this new set is closely related to spherical polar coordinates:
x = r sin φ cos θ,
y = r sin φ sin θ,
z = r cos φ.
The metric in these coordinates is
ds2 = dr2 + r2dφ2 + r2 sin2 φdθ2.
The invariants associated with these coordinates are defined as [17]
R1 =
r
= norm(D),
2 sin φ = r 3
R2 = q 3
R3 = − cos 3θ = 3√6
2
,
norm( D)
norm(D)
det( D)
norm( D)3
(14)
(15)
(16)
,
where, using the definition of the norms in Eq. (11), one can see that R2 is equivalent to the usual
definition of fractional anisotropy (see Eq. (1)).
The structure of these spherical invariants is very similar to the cylindrical invariants. R1, like
K1, is a measure of the magnitude of the diffusion tensor but is now the radial coordinate, r, which
corresponds to the norm of the diffusion tensor. Again, R2 and R3 measure the anisotropy of the
tensor. R2, as we have already mentioned, is exactly equivalent to fractional anisotropy, Af , and is
related to the polar angle, φ. It is also similar to K2 in its geometrical interpretation in that it is also
(proportional to) the perpendicular distance from a point (r, φ, θ) to the z-axis (φ = 0 direction) but
normalised by the radius, r. R3 is exactly the same as K3, the mode, and thus quantifies the diffusion
ellipsoid shape. As with the cylindrical coordinates, it is also related to the azimuthal angle, θ.
As with the cylindrical set, we can regard {Ri} as coordinates on a related manifold R and
demonstrate the orthogonality of these invariants by showing that the metric is purely diagonal;
ds2 = dR 2
1 +
3
1
R 2
2 − R 2
2
dR 2
2 +
2
1 R 2
2R 2
27(1 − R 2
3 )
dR 2
3 .
(17)
Log-Euclidean Coordinates and Invariants
Recently in the literature there has been some debate over the way in which diffusion tensors should
be averaged and interpolated [19]. The simplest way to average tensors is to treat them as one would
treat real numbers by directly adding them and dividing by the number of tensors. A consequence of
this (Euclidean-averaging) method is that if the original tensors had the same traces then the resulting
tensor also has this trace. There is then a natural affinity between this method and the invariant set
PLoS ONE
7
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
{Ki} since K1 (the trace) is preserved in such circumstances. This is because the condition that K1 is
constant defines a flat plane in the eigenvalue space (with its assumed Euclidean metric) and geodesics
between points in this plane are straight lines and therefore will remain in the plane.
The set {Ri} is also consistent with Euclidean averaging even though none of the set members are
generally preserved in the sense that K1 is. This is because this set is also derived from the assumption
of a Euclidean metric on the eigenvalue space. However, if one were interested in preserving R1 in the
same way that K1 is preserved in the cylindrical set, then one would need to average over the surface
of a sphere. As far as we are aware, no such averaging method has been proposed in the literature
but one such method would be to simply average the square of the diffusion tensors and find the
appropriate square-root. This method would preserve the tensor norm for tensors of equal norm.
An alternative method of averaging which has been discussed and explored in the literature is the
log-Euclidean method [9,10,20]. This is similar to the Euclidean method except that the logarithm of
the tensors is averaged. This method preserves the tensor determinant. It would therefore be natural
to ask whether there exists an invariant set {Li} such that L1 is related to the determinant. Here we
show that such a set does exist and had already been found by Criscione et al. [18].
We begin with Eq. (6) but insert the map
ℓκ : (λ1, λ2, λ3) 7→ (ln κλ1, ln κλ2, ln κλ3),
(18)
where κ > 0 is an arbitrary real parameter with dimensions s/m2, prior to the application of the
rotation matrix. ℓκ maps the octant of eigenvalue space to the full Euclidean space. We now assume
that the metric of Eq. (7) is valid. This is now our starting point, not the metric of Eq. (5), which is
no longer valid.
We now follow the construction of the cylindrical invariant set equations (8) to (10) except that
now we have
L1 =
L2 =
√3z
ρ
= tr(Λ),
= norm(Λ),
(19)
L3 = − cos 3θ = 3√6
3 tr(Λ)I. By using the fact that tr(Λ) = ln(κ3 det(D)) and Λ =
det(Λ)
norm(Λ)3
,
where Λ = ln κD and Λ = Λ − 1
ln(cid:2)D/ det(D)1/3(cid:3) we can re-express the above as
L2 = vuut
3
L1 = ln(κ3D),
L3 =
,
D1/3(cid:19)(cid:21)2
Xi=1 (cid:20)ln(cid:18) λi
3√6Q3
i=1 ln(cid:18) λi
D1/3(cid:19)
L 3
2
(20)
,
where λi are the eigenvalues of D and D = det(D) = λ1λ2λ3.
As we remarked earlier, the set of invariants {Li} is identical to that found in the paper by
Criscione et al. [18] (except for the inclusion of the parameter κ). The cylindrical invariants {Ki} of
PLoS ONE
8
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
Ennis and Kindlmann [17] were derived from the invariants of Criscione et al. but the latter authors'
log-Euclidean invariants were not mentioned in the former authors' paper. Furthermore, L2 was
discovered independently by Batchelor et al. [9] and Fletcher and Joshi [10] from their investigations
of the space of diffusion tensors. Both sets of authors referred to L2 as geodesic anisotropy.
The structure of these log-Euclidean invariants is similar to the cylindrical and spherical sets.
L1 ∈ (−∞,∞) is the measure of overall magnitude of the diffusion tensor. L2 ∈ [0,∞) is similar
to fractional anisotropy in these log-Euclidean coordinates and has a similar interpretation, being
the perpendicular distance, ρ, between the point (z, ρ, θ) and the z-axis (isotropy). Since this is the
distance of a path which is also a geodesic in this space (a straight line in Euclidean space), L2 was
called geodesic anisotropy [9, 10]. The same appellation could equally be applied to K2, however. A
difference worth noting between K2 and L2 is that the latter is scale-independent, as is evidenced by
the absence of the parameter κ.
L3 ∈ [−1, 1] plays an identical role to the mode as defined by K3 (and R3). However, since we
have now assumed a different metric on the eigenvalue space, the parameterised description of the
four archetypal cases is now different to that given in (2) and can be given by
(cid:8)D1/3, D1/3, D1/3(cid:9) ,
(cid:8)D1/3eα, D1/3, D1/3e−α(cid:9) ,
(cid:8)D1/3e2α, D1/3e−α, D1/3e−α(cid:9) , Prolate
(cid:8)D1/3eα, D1/3eα, D1/3e−2α(cid:9) ,
Oblate
Isotropic
Orthotropic
(21)
{λ1, λ2, λ3} =
where α ∈ (0,∞). Alternatively, as we did previously in (12), the prolate and oblate cases can be
parameterized together as
{λ1, λ2, λ3} = nD1/3e2β, D1/3e−β, D1/3e−βo ,
(22)
where β ∈ (−∞,∞) can be broken into a union of three intervals as β ∈ (−∞, 0)∪ [0] ∪ (0,∞). These
first and last intervals correspond to the oblate and prolate cases respectively and the point β = 0 is
the isotropic case. Using this parameterisation, L3 = sgn(β). As with K3, a discontinuity exists as
β passes from negative to positive and L3 at β = 0 (isotropy) is undefined. For the orthotropic case,
L3 = 0 for all α including the isotropic case, i.e. α ∈ [0,∞).
The orthogonality of these invariants is guaranteed by the fact that they have exactly the same
relation to the cylindrical coordinates as the {Ki}. Therefore, the metric is exactly analogous to
Eq. (13).
Conformal Maps, Curvilinear Coordinates and Invariants
In the previous sections, the orthogonal invariant sets have a common structure; the first invariant
measures the magnitude of the diffusion tensor; the second invariant measures the magnitude of the
anisotropy and the third invariant provides a measure of the shape of the ellipsoid. In this final section,
we will describe an invariant set with a slightly different structure.
We begin by recalling that the invariant sets {Li} and {Ki} were based on an underlying cylindrical
coordinate system {z, ρ, θ} and that the second and third members of these sets are related to the
plane of the polar coordinates {ρ, θ}. This plane is naturally regarded as the two-dimensional space
PLoS ONE
9
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
of real numbers R2 but can equally be viewed as the complex plane C and, in this context, there are
a family of transformations which preserve angles; the conformal maps [21].
We therefore seek conformal transformations of these coordinates which produce alternative in-
variant measures {Ci} (other than those in Eqs. (10) or (19)) such that C1 = K1 or L1.
Defining the complex number Z = x + ıy = ρeıθ, we seek conformal maps F : C → C such
that W = F (Z) = U (x, y) + ıV (x, y). Such conformal maps are complex analytic functions with non-
vanishing complex derivative (except at critical points) and guarantee the preservation of orthogonality
via the Cauchy-Riemann equations
∂U
∂x
=
∂V
∂y
and
∂U
∂y
∂V
∂x
.
= −
(23)
Potential invariants can then be constructed from (C2, C3) = (f2(U ), f3(V )), where f2 and f3 are
differentiable functions.
The simplest maps to provide invariants appear to be Fn : Z 7→ Z 3n for integers n ≥ 1, with more
complicated maps being constructed from convergent polynomials or fractional linear transformations
thereof. Using the simplest, non-trivial map F1, with an additional π/2 rotation, we make the following
assignments;
ρ3 sin 3θ,
C2 =
C3 = −ρ3 cos 3θ.
(24)
At this stage, the invariants C2 and C3 are expressed in terms of polar coordinates and we could
choose to relate them to {Ki} or {Li}. From this point on, we will relate them to the log-Euclidean
set since the resulting invariants will then be scale-independent.
The rotation used in obtaining the invariants above conveniently introduces the minus sign for C3
3 . These two
so that, comparing it to the mode L3 (Eq. (19)), we see that C3 ∝ L3 and C2 ∝ p1 − L 2
invariants can therefore be re-expressed as
(25)
(26)
C2 = L 3
C3 = L 3
2 p1 − L 2
2 L3.
3 ,
The metric for coordinates {Ci} on the related manifold C is
2 + dC 2
3 (cid:1)2/3 .
2 + C 2
9(cid:0)C 2
ds2 =
dC 2
1 +
1
3
dC 2
3
The invariants {Ci} therefore form an orthogonal set and, as mentioned above, these invariants possess
a slightly different structure to the previous three sets {Ki}, {Ri}, and {Li}. In the previous sets,
the second invariant (i = 2) measures the magnitude of anisotropy whereas the third invariant (i = 3)
measures the shape of the ellipsoid. As shown above in Eq. (25), C2 and C3 are both functions of L2
and L3; both contain information about the magnitude of the anisotropy and the ellipsoid shape but
C2 and C3 contain different information about the shape. Table 1 shows a comparison between the
spherical measures, R2 and R3, and the curvilinear measures, C2 and C3. It can be seen that C3 is
similar to R3 except that the magnitude of C3 is modulated by a positive function of L2. Therefore,
C3 measures the degree of prolateness or oblateness of the ellipsoid, the two cases being differentiated
by the sign of C3. C2, on the other hand, measures the degree of orthotropy. This is in contrast to
PLoS ONE
10
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
Table 1. Spherical and curvilinear anisotropies
Isotropic
Orthotropic
C2
Ellipsoid shape R2 R3
0
0
0
> 0
0 > 0
> 0
1
> 0 −1
Prolate
Oblate
C3
0
0
0 > 0
0 < 0
Comparison of the spherical (R2, R3) and curvilinear measures (C2, C3) of anisotropy for the four
archetypical ellipsoid shapes.
the spherical measures in which orthotropy is inferred from the absence of prolateness or oblateness;
R3 = 0 but R2 6= 0.
We also note that the discontinuity in L3 and R3 as one passes through the isotropic state is absent
from C2 and C3. If the parameterisation of Eq. (22) is used for the prolate and oblate cases, the pair
(C2, C3) = (0, 63/2β3) and it is clear that smooth behaviour is exhibited for all values of β. Likewise,
using the parameterisation for the orthotropic case in Eq. (21), (C2, C3) = (23/2α3, 0), which is also
smooth (up to second derivative) for all values of α, even if we extend the parameterisation to negative
values; α ∈ (−∞,∞).
To calculate the curvilinear measures, we can first calculate the log-Euclidean measures using
Eq. (19) or (20) and use (25) together with C1 = L1, but it is also interesting and useful to show how
they can be calculated directly from the diffusion and deviatoric tensors;
C1 = tr(Λ),
C2 = qnorm(Λ)6 − 54 det(Λ)2,
C3 = 3√6 det(Λ).
(27)
Finally, we note that in constructing the curvilinear set of invariants {Ci}, we have done so using
the invariant sets related to cylindrical coordinates on the eigenvalue space. This was convenient
because the polar coordinates define a plane on which we could employ the conformal transformations.
However, something similar can be done with the spherical set {Ri}. In this case, a stereographic
projection can be used to conformally map the spherical coordinates {φ, θ} to the complex plane and
then invariants can be defined similarly to those in this section.
Discussion
Up to permutations of the eigenvalues, any invariant set can be inverted to return the eigenvalues
(see Criscione et al. [18]). Therefore any two invariant sets contain the same information. However,
different sets display this information in different ways and can make certain aspects of this information
less or more clear.
As described in the Theory section, the sets {Ki}, {Ri} and {Li} are all similar in structure. What
distinguishes the set {Ci} is that it makes orthotropic anisotropy more explicit. In the previous three
sets, orthotropy is inferred from the combination of, for example, R2 > 0 and R3 = 0, whereas C2 is
PLoS ONE
11
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
a more direct measure of the degree of orthotropy. That is the single invariant C2 provides a much
clearer picture of this aspect of the anisotropy.
Whilst it could be seen as a disadvantage that the invariants C2 and C3 have a cubic non-linearity
(see equation (25)), it is this property that has removed the discontinuity of the mode at isotropy.
This feature of C2 and C3 may prove to be advantageous when dealing with noisy diffusion tensors
that are close to isotropic. Since the mode, R3, possesses a discontinuity at isotropy, small fluctuations
in the eigenvalues can produce large fluctuations in R3. On the other hand, since C2 and C3 behave
smoothly at isotropy it is expected that the effects of noise will be decreased in the region of isotropic
tensors.
Limitations of the Study, Open Questions and Future Work
In the current work, we have focused on the mathematical development of the curvilinear invariants,
their properties and their relation to other invariants. Our current interest is in the application of
the new invariant set to DTI measurements on articular cartilage in the hope that it will aid in
understanding the nature of the collagen architecture which might exhibit a depth-dependence. Our
initial investigations suggest that our invariants show there is a difference in the structure between the
superficial zone (oblateness) and the deep zone (prolateness) of articular cartilage. We are currently
in the process of preparing this data for publication.
What also remains to be verified is the claim that the curvilinear invariants possess better noise
tolerance for approximately isotropic tensors. This could be achieved, for example, via the statistics
of computer simulations of noisy tensors.
Conclusions
We have explored the cylindrical and spherical invariant sets of the diffusion tensor that were proposed
by Ennis and Kindlmann [17], showing how they relate to the geometry of the eigenvalue space and
providing an alternative method for the proof of their orthogonality (via the metric). We have also
shown how a log-Euclidean set is closely related to the cylindrical set and how this and the other sets
are consistent with various averaging schemes.
A curvilinear invariant set {Ci} was developed which has a different structure to the preceding
three sets, making orthotropy more explicit, and which is expected to be more suitable for measuring
low degrees of anisotropy and isotropy. In addition to the usual analysis of eigenvectors and fractional
anisotropy from DTI studies on tissues such as articular cartilage, the curvilinear invariants may
provide an improved assessment of morphology and function of cartilage and other biological tissues.
References
1. Le Bihan D, Johansen-Berg H (2012) Diffusion MRI at 25: Exploring brain tissue structure and
function. Neuroimage 61: 324 -- 341.
PLoS ONE
12
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
2. Sinha U, Sinha S, Hodgson JA, Edgerton RV (2011) Human soleus muscle architecture at
different ankle joint angles from magnetic resonance diffusion tensor imaging. Journal of Applied
Physiology 110: 807 -- 819.
3. Benson AP, Bernus O, Dierckx H, Gilbert SH, Greenwood JP, et al. (2011) Construction and val-
idation of anisotropic and orthotropic ventricular geometries for quantitative predictive cardiac
electrophysiology. Interface Focus 1: 101 -- 116.
4. de Visser SK, Crawford RW, Pope JM (2008) Structural adaptations in compressed articular
cartilage measured by diffusion tensor imaging. Osteoarthritis and Cartilage 16: 83 -- 89.
5. Filidoro L, Dietrich O, Weber J, Rauch E, Oerther T, et al. (2005) High-resolution diffusion
tensor imaging of human patellar cartilage: Feasibility and preliminary findings. Magnetic Res-
onance in Medicine 53: 993 -- 998.
6. Meder R, de Visser SK, Bowden JC, Bostrom T, Pope JM (2006) Diffusion tensor imaging
of articular cartilage as a measure of tissue microstructure. Osteoarthritis and Cartilage 14:
875 -- 881.
7. Le Bihan D, Mangin JF, Poupon C, Clark CA, Pappata S, et al. (2001) Diffusion tensor imaging:
Concepts and applications. Journal of Magnetic Resonance Imaging 13: 534 -- 546.
8. Basser PJ, Pierpaoli C (1996) Microstructural and physiological features of tissues elucidated
by quantitative-diffusion-tensor MRI. Journal of Magnetic Resonance Series B 111: 209 -- 219.
9. Batchelor PG, Moakher M, Atkinson D, Calamante F, Connelly A (2005) A rigorous framework
for diffusion tensor calculus. Magnetic Resonance in Medicine 53: 221 -- 225.
10. Fletcher PT, Joshi S (2007) Riemannian geometry for the statistical analysis of diffusion tensor
data. Signal Processing 87: 250 -- 262.
11. Bahn MM (1999) Invariant and orthonormal scalar measures derived from magnetic resonance
diffusion tensor imaging. Journal of Magnetic Resonance 141: 68 -- 77.
12. Hasan KM, Alexander AL, Narayana PA (2004) Does fractional anisotropy have better noise im-
munity characteristics than relative anisotropy in diffusion tensor MRI? An analytical approach.
Magnetic Resonance in Medicine 51: 413 -- 417.
13. Peled S, Gudbjartsson H, Westin CF, Kikinis R, Jolesz FA (1998) Magnetic resonance imaging
shows orientation and asymmetry of white matter fiber tracts. Brain Research 780: 27 -- 33.
14. Westin CF, Maier SE, Mamata H, Nabavi A, Jolesz FA, et al. (2002) Processing and visualization
for diffusion tensor MRI. Medical Image Analysis 6: 93 -- 108.
15. Basser PJ, Pajevic S (2000) Statistical artifacts in diffusion tensor MRI (DT-MRI) caused by
background noise. Magnetic Resonance in Medicine 44: 41 -- 50.
16. Pierpaoli C, Basser PJ (1996) Toward a quantitative assessment of diffusion anisotropy. Mag-
netic Resonance in Medicine 36: 893 -- 906.
PLoS ONE
13
2013 Volume 8 Issue 11 e78798
Invariant Sets of the Diffusion Tensor
17. Ennis DB, Kindlmann G (2006) Orthogonal tensor invariants and the analysis of diffusion tensor
magnetic resonance images. Magnetic Resonance in Medicine 55: 136 -- 146.
18. Criscione JC, Humphrey JD, Douglas AS, Hunter WC (2000) An invariant basis for natural
strain which yields orthogonal stress response terms in isotropic hyperelasticity. Journal of the
Mechanics and Physics of Solids 48: 2445 -- 2465.
19. Pasternak O, Sochen N, Basser PJ (2010) The effect of metric selection on the analysis of
diffusion tensor MRI data. Neuroimage 49: 2190 -- 2204.
20. Arsigny V, Fillard P, Pennec X, Ayache N (2006) Log-euclidean metrics for fast and simple
calculus on diffusion tensors. Magnetic Resonance in Medicine 56: 411 -- 421.
21. Jeffrey A (1986). Mathematics for Engineers and Scientists. Third ed. Berkshire, England: Van
Nostrand Reinhold (UK). pp. 444 -- 455.
PLoS ONE
14
2013 Volume 8 Issue 11 e78798
|
1808.09764 | 1 | 1808 | 2018-08-29T12:40:02 | Is Actin Filament and Microtubule Growth Reaction- or Diffusion-Limited? | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.SC"
] | Inside cells of living organisms, actin filaments and microtubules self-assemble and dissemble dynamically by incorporating actin or tubulin from the cell plasma or releasing it into their tips' surroundings. Such reaction-diffusion systems can show diffusion- or reaction-limited behaviour. However, neither limit explains the experimental data: while the offset of the linear relation between growth speed and bulk tubulin density contradicts the diffusion limit, the surprisingly large variance of the growth speed rejects a pure reaction limit. In this Letter, we accommodate both limits and use a Doi-Peliti field-theory model to estimate how diffusive transport is perturbing the chemical reactions at the filament tip. Furthermore, a crossover bulk density is predicted at which the limiting process changes from chemical reactions to diffusive transport. In addition, we explain and estimate larger variances of the growth speed. | physics.bio-ph | physics |
Is Actin Filament and Microtubule Growth Reaction- or
Diffusion-Limited?
Johannes Pausch*, Gunnar Pruessner
Department of Mathematics, Imperial College London, UK
August 30, 2018
Abstract
Inside cells of living organisms, actin filaments and microtubules self-assemble and dissemble dynamically by incorpo-
rating actin or tubulin from the cell plasma or releasing it into their tips' surroundings. Such reaction-diffusion systems can
show diffusion- or reaction-limited behaviour. However, neither limit explains the experimental data: while the offset of the
linear relation between growth speed and bulk tubulin density contradicts the diffusion limit, the surprisingly large variance
of the growth speed rejects a pure reaction limit. In this Letter, we accommodate both limits and use a Doi-Peliti field-theory
model to estimate how diffusive transport is perturbing the chemical reactions at the filament tip. Furthermore, a crossover
bulk density is predicted at which the limiting process changes from chemical reactions to diffusive transport. In addition,
we explain and estimate larger variances of the growth speed.
Microtubules and actin filaments are structures that poly-
merise by incorporating and releasing their diffusively mov-
ing building blocks, tubulin and actin. They are responsi-
ble for growth, shape, movement, and transport processes,
among others, and can span the entire cell [3]. Their dynam-
ics have been studied intensively experimentally1 and the-
oretically.2 However, many questions about their dynamics
remain debated or completely unanswered. This includes:
Is their growth limited by diffusion to their tip [18, 8] or by
the chemical reaction rates for incorporation [1, 8]? How
can the large variance of their growth be explained [12]?
In experiments, the filament growth speed (cid:104)v(cid:105) can be
measured as a function of the bulk density ζ of tubulin or
actin. An effective incorporation coefficient kon and effec-
tive release rate koff are determined as parameters of a linear
fit of growth speed data over bulk concentration ζ
(cid:104)v(cid:105) = h(konζ − koff),
(1)
where h is the effective growth length per incorporated par-
ticle. There are two mean field approaches for modeling the
filament growth speed (cid:104)v(cid:105).
The first approach assumes that diffusive transport is
quicker than the chemical reactions [16, 4, 6, 25, 1] and
that polymerisation is therefore reaction-limited. This ef-
fectively infinite diffusivity D implies that particle concen-
tration is homogeneous, and in particular, that there is no
significant depletion close to the filament tip. Thus, fila-
ment growth is determined by two Poisson processes, incor-
poration with rate λζ and release with rate τ, each of which
is associated with a step of length h. The two competing
*corresponding author: [email protected]
1See Table S1 in the Supplemental Information of [12] for microtubule
assembly rates. For actin filaments, assembly rates can be found in [20,
10, 15, 11].
2Theoretical work includes [16, 9, 4, 6, 25, 1, 18] for microtubules and
[9, 10, 14] for actin filaments.
1
processes create a Skellam distribution [21] with expected
growth speed (cid:104)v(cid:105)R and effective diffusion constant Deff
and Deff = h2(λζ + τ ).
(2)
(cid:104)v(cid:105)R = h(λζ − τ )
However, in comparison with experiments [12, 10, 15], Deff
is too small. Furthermore, it implies an independence of the
growth speed from the viscosity of the medium, which was
rejected experimentally for actin filaments [8] and micro-
tubules [26]. Thus, a purely reaction-limited behaviour is
rejected.
The second approach assumes that transport by diffusion
is slower than the chemical reactions [8, 18] and that self-
assembly is therefore diffusion-limited. This implies that,
due to the incorporation into the tip, the building blocks are
depleted locally. The growth speed is determined by the
diffusive flux to the tip. If the protein concentration c(x, t)
follows a steady state diffusion equation 0 = D∆c(x, t),
the particle flux J through the absorbent reaction sphere
of radius R and the growth length h determine the growth
speed (cid:104)v(cid:105)D, which is equivalent to Smoluchowski coagula-
tion [24]
(cid:104)v(cid:105)D = Jh = 4πDRhζ.
(3)
However, this limit cannot accommodate a release rate as
any released particle would immediately be reabsorbed be-
fore diffusion can transport it away from the reaction sur-
face. Furthermore, using this approach with typical pa-
rameters of microtubule assembly, only a small reduction
of the bulk tubulin density to 89% is found at the reac-
tion surface [18]. It therefore is not completely absorbent,
as is theoretically suggested in [5]. According to the
Stokes-Einstein equation [22] for small Reynolds numbers,
diffusion-limited growth implies that viscosity and incor-
poration rate are inversely proportional [2] without offset.
Tested in [8] and [26], small offsets for the growth of micro-
tubules and the barbed actin filament ends are found, while
Figure 1: Filament growth speed (cid:104)v(cid:105) as function of the bulk
density ζ of its building blocks. The diffusion-limited (cid:104)v(cid:105)D
(red), the reaction-limited (cid:104)v(cid:105)R (blue), and the real growth
speed (cid:104)v(cid:105) (black) are shown. A: The system is reaction-
limited, B: Reaction-limited growth changes to diffusion-
limited growth at the crossover bulk density ζ×. Local
densities at the tip can be read off as the bulk density at
which reaction-limited growth leads to the same speed (dot-
ted line).
a significant offset is found for the pointed ends of actin fila-
ments. Thus, filament growth cannot be perfectly diffusion-
limited either. Furthermore, this model is not probabilistic
and therefore, it is not clear how to derive a variance of the
growth speed.
The diffusion- (Eq. (3)) and reaction-limited (Eq. (2)), as
well as the measured growth speed are schematically de-
picted in Fig. 1. Measured growth and shrinking speeds
will always be slower than both limits. When shrinking, the
chemically possible shrinking speed is not attained as dif-
fusion slows down transport away from the tip, leading to
a locally higher particle density. It can be read off the plot
as the density at which the reaction-limited is equal to the
measured speed (dotted line). Analogously, when growing,
the reaction-limited speed is not attained either because dif-
fusion fails to maintain the bulk particle density around the
tip. The local density at the tip can be read off again as
the density at which the reaction-limited speed is equal to
the measured one (dotted line). In principle, there are three
cases: the purely reaction limited case A, the mixed case
B, where a crossover from reaction-limited to diffusion-
limited behaviour occurs at the crossover bulk density ζ×,
and case C (not shown), where shrinking does not occur and
the growth is limited by diffusion at all particle densities.
Theoretically, progress can be made by going beyond
mean field theory which allows to calculate how chem-
ical reactions are perturbed by diffusive transport of the
reactants. Here, filament self-assembly is modelled on a
three-dimensional lattice and described by a master equa-
tion (Supplementary Material, Eq. A2). Following work by
Doi [7] and Peliti [19], this model is transformed into a field
theory. The derivation of the field theory is outlined briefly
in the Supplementary Material, in Sections A2 and A3.
In our model, the building blocks and the filament tip
are represented by fields, which are interpreted as time-
dependent probability distributions of their positions. Due
Figure 2: Schematic of the filament self-assembly process.
Its building blocks move diffusively in R3. The filament tip
is on a lattice with spacing h. Particles can be incorporated
in the filament (coefficient λ) and released by it (rate τ).
to the field representation, they do not have a size. How-
ever, the finite size of the proteins is an important element
of the step-wise filament growth. Therefore, our model is
set up on a hybrid, three dimensional space: While parti-
cles move in continuous space R3, the filament tip is re-
stricted to a discrete line Z with lattice constant h, over-
laying the z-axis in R3, see Fig. 2. The persistence length
P and flexural rigidity K of the filament we model are thus
effectively infinite. This is a good approximation for micro-
tubules (P ≈ 5200µm, K ≈ 2·10−23N m2), while for actin
filaments, this approximation is slightly worse (P ≈ 18µm,
K ≈ 7 · 10−26N m2) [13]. As we are modelling only a sin-
gle polymer instead of the 13 microtubule protofilaments or
the 2 actin filament strands, we interpret the lattice spacing
h as the effective growth step length.
Each of the fields exists as an annihilation field and a
creation field: ϕ(x, t) and ϕ†(x, t) for the building blocks,
as well as ψj(t) and ψ†j (t) for the filament tip. The creation
field initiates a single particle or filament tip at the specified
position and time, whereas the annihilation field measures
their number at the point stated. Creation fields often appear
as Doi-shifted fields [7], e.g. ϕ†(x, t) = (cid:101)ϕ(x, t) + 1. In
addition, the particle annihilation field is shifted to measure
deviations from the bulk density ζ, i.e. ϕ(x, t) = ϕ(x, t) +
ζ. Between creation and annihilation, the system evolves
by the stochastic processes included in the model.
There are six microscopic processes in our model. The
units of the corresponding coefficients, denoted by [. . . ],
are written as monomials of T (time) and L (length).
• particle diffusion with constant D, [D] = T −1L2
• actin or tubulin absorption by the filament tip with co-
efficient λ and subsequent movement of the tip by dis-
tance h in the +z direction, [λ] = T −1L3
• particle release from the filament tip with rate τ and
subsequent movement of the tip by distance h in the
−z direction, [τ ] = T −1
• actin/tubulin creation with coefficient γ,
[γ] =
2
ζhvihviDhviRhvi4πDRhhλhkonhτhkoff0AζhvihviDhviRhvi4πDRhhλhkonζ×hτhkoff0BλτhT −1L−3
corporation or release of tubulin. It is described by
• extinction of actin/tubulin with rate r, [r] = T −1
• extinction of the filament tip with rate , [] = T −1
Alin-F-stat =
However, due to incorporation and release, the bilinear
part includes jumps in steps of 1z
Creation and extinction of the building blocks is balanced
such that a constant bulk density ζ = γ/r is created. The
two extinction processes are included in the field theory to
enforce causality. After calculations, we let parameters γ, r
and tend to zero while keeping the ratio γ/r = ζ constant.
Thus, the spontaneous extinction and creation are removed
while a bulk density remains included.
All of the processes above are reflected in the action func-
tional A of our model which splits up into a bilinear part
and an interaction part A = Alin + Aint. The diffusion and
extinction of particles is represented in the particle bilinear
part
(cid:101)ϕ(x, t)(cid:0)
−∂t + D∆ − r(cid:1) ϕ(x, t)d3xdt,
Alin-P =
(4)
(cid:90)
R4
where ∆ is the spatial Laplace operator.
(cid:90)
(cid:88)
R
j∈Z
(cid:90)
(cid:88)
(cid:16)
R
j∈Z
Alin-F-mov =
(cid:101)ψj(t)(cid:0)
−∂t − (cid:1)ψj(t)dt.
(5)
growing
(cid:125)(cid:124)
(cid:123)
(cid:122)
λζ(cid:0)(cid:101)ψj+1z − (cid:101)ψj
(cid:1)ψj
+ τ(cid:0)(cid:101)ψj−1z − (cid:101)ψj
(cid:1)ψj
(cid:123)(cid:122)
(cid:124)
shrinking
(6)
(cid:17)
(cid:125)
dt,
where we omitted the time dependence of the fields for
better readability. The first part corresponds to filament
growth, while the second part describes shrinking of the fil-
ament. Jumps on the lattice are indicated by ±1z.
Alin-P + Alin-F-stat + Alin-F-mov.
All three bilinear actions together make up Alin =
The filament tip is stationary without the processes of in-
(cid:90)
(cid:20)
(cid:17)
(cid:16)(cid:0)(cid:101)ψj+1z − (cid:101)ψj
(cid:88)
(cid:1)ψj ϕ(hj)
The interaction part of A has the form
−(cid:101)ψjψj(cid:101)ϕ(hj) ϕ(hj)
− ψj(cid:101)ϕ(hj) ϕ(hj)
(cid:124)
(cid:125)
(cid:123)(cid:122)
(cid:124)
(cid:125)
(cid:123)(cid:122)
(cid:125)
(cid:123)(cid:122)
(cid:124)
(cid:21)
(cid:1)ψj(cid:101)ϕ(hj)
+(cid:0)τ(cid:101)ψj−1z − λζ(cid:101)ψj
+ (τ − λζ)(cid:101)ϕ(hj)ψj
(cid:125)
(cid:123)(cid:122)
(cid:124)
(cid:123)(cid:122)
(cid:124)
(cid:125)
Aint =
j∈Z
dt,
(a)
(c)
(b)
λ
R
(d)
(e)
(7)
The interplay between propagation and interaction in a
system governed by the action A can be calculated using
the path integral. The system may be initialised by placing
a filament tip at position hj0 = 0 at time t0 = 0. Then,
the system evolves and particle concentrations, the filament
tip positions, or moments of their distributions can be mea-
sured at a later point in time. In general, if the observable
where the time dependency is omitted for better readability.
The different parts of the interaction describe the follow-
ing processes:
(a) the filament grows, i.e.
the filament tip moves one
step in the positive z direction upon incorporating an
actin/tubulin particle;
(b) as particles are incorporated into the tip, its density is
reduced locally, resulting in anticorrelations of the tip
and the particle density;
(c) particle density is reduced by incorporation into the
tip;
(d) in the presence of a tip, the particle density is increased
by spontaneous release (τ) and decreased by incorpo-
ration (λζ), leading to corresponding correlations of
tip and particle densities;
(e) τ: particle density is increased because the filament
releases a particle; λζ: particle density is decreased
because the filament incorporates a particle.
and
propagations
Field-theoretic
are
schematically represented by Feynman diagrams. Particle
propagation is drawn as a straight red line, filament propa-
gation is depicted as a curly blue line, and interactions are
illustrated as vertices, see Fig. 3.
interactions
Figure 3: The stochastic processes that appear in Aint,
Eq. (7), are represented as amputated vertices in Feyn-
man diagrams. Curly blue lines represent filaments, while
straight red lines stand for their building blocks, actin or
tubulin. All Feynman diagrams are read from right to left.
3
λψeψϕ(a)−λψeψeϕϕ(b)−λϕeϕψ(c)τ−λζψeψeϕ(d)τ−λζψeϕ(e)The loop correction term takes into account that, due to
diffusive transport, the reaction-limited speed is reduced
further because the locally depleted particles have to reach
the reaction sphere of radius R. Considering the first term
of the loop correction, if the diffusion is strong, the chem-
ical reactions are less hindered by slow transport and the
growth speed is closer to its reaction limit. This diffusion
correction is itself corrected in the second term of the loop
correction which describes how quickly the tip reaches re-
gions that are less depleted. It predicts a non-linear depen-
dence of the growth speed (cid:104)v(cid:105) on the bulk particle density ζ,
which so far has not been observed experimentally. There-
and ignore the latter in
fore, we assume
the following. Thus, given the diffusion-limited growth co-
efficient 4πDR and the observed kon and koff, Eq. (1), we
can calculate the reaction-limited λ as well as the offset τ
4πDR (cid:29) hλζ−τ
8πD2
1
λ = kon
4πDR
4πDR − kon
and
τ = koff
4πDR
4πDR − kon
,
(12)
to the results in [17] for general
which is equivalent
reaction-diffusion systems.
It follows, that the observed
growth coefficient kon is smaller than the reaction- and the
λ < 1 and
diffusion-limited growth coefficients, i.e.
4πDR < 1, as shown in Fig. 1.
Furthermore, there is a bulk particle density ζ× at which
diffusion becomes the defining limitation in comparison to
reaction (Fig. 1, B):
kon
kon
ζ× =
koff
2kon − 4πDR
.
(13)
For koff = 0, if 4πDR > 2kon, then growth is reaction-
limited otherwise it is diffusion-limited.
One of the open questions for microtubule and actin fila-
ment growth is the origin of large fluctuations [12, 10]. In
part, they can be explained as an overlap of the fluctuations
of the reaction and diffusion processes. This is quantified
by the effective diffusion constant (derived in SI, Sec. A6)
(cid:18)
(cid:19)
Deff,2 =h2(λζ + τ )
λ
4πDR
1 −
+ λ
hλζ − τ
2Dπ
,
(14)
which is larger than the reaction-limited effective diffusion
in Eq. (2).
In conclusion, the field theoretic model allows us to cal-
culate how the reaction limit of microtubule and actin fila-
ment growth is undermined due to imperfect diffusive sup-
ply of tubulin or actin. It also predicts larger growth fluctua-
tions compared to a model purely based on reactions. Given
the overlap of fluctuations due to chemical reactions and
diffusive transport, it is likely for the filament growth speed
to exhibit correlations in time, the study of which would be
compelling for future research.
The authors thank Robert Endres, Guillaume Salbreux
and Thomas Surrey for very helpful discussions.
that we want to measure is represented by a combination
of fields O(t), then its time-dependent, spatial probability
distribution is given by the following path integral (see e.g.
[23] for a detailed review)
(cid:90)
∞(cid:88)
(cid:104)O(t)ψ†0(0)(cid:105):=
(cid:96)=0
D[ϕ, ψ]O(t)ψ†0(0)e−Aprop (−Aint)(cid:96)
(cid:96)!
.
(8)
(cid:90)
∞(cid:88)
(cid:96)=0
Formally, this integral is summing all variations of all fields
involved of all stochastic processes possibly occurring. The
path integral is normalised such that
(cid:104)1(cid:105) =
D[ϕ, ψ]e−Aprop (−Aint)(cid:96)
(cid:96)!
= 1.
(9)
The (cid:96)-th term of the series is the contribution of all pro-
cesses with (cid:96) interactions.3 The distribution of the fila-
ment's tip position j is given by (cid:104)ψj(t)ψ†0(0)(cid:105). The ex-
pected growth speed and its variance are then determined
by calculating the filament's expected position and its vari-
ance after time t. In the following, we consider three ap-
proximations of (cid:104)ψj(t)ψ†0(0)(cid:105): Firstly, (cid:104)ψj(t)ψ†0(0)(cid:105)0 is the
reaction-limited distribution, cutting the sum in Eq. (8) at
(cid:96) = 0, which results in only one Feynman diagram shown
in Eq. (10a). Secondly, (cid:104)ψj(t)ψ†0(0)(cid:105)2 is terminating the
sum at (cid:96) = 2, resulting in two diagrams shown in Eq. (10a)
and (10b). Thirdly, (cid:104)ψj(t)ψ†0(0)(cid:105)Dy is the Dyson sum,
which contains terms of all orders but selects only those
Feynman diagrams whose loops are arranged daisy-chain-
like, shown in Eq. (10):
(cid:104)ψj(t)(cid:101)ψ0(0)(cid:105)Dy
=
+
+
(10a)
(10b)
+ . . .
(10c)
A priori, it is not clear which truncation of the path integral
is a good approximation of the observable. However, a good
agreement of the approximate result with experimental data
indicates that the processes which were not included in the
calculation rarely occur under experimental conditions.
The second and third approximation for the average fila-
ment growth speed are
(cid:18)
(cid:16)
(cid:124)
(cid:104)v(cid:105)2 =h(λζ − τ )
1 − λ
4πDR −
1
hλζ − τ
8πD2
Loop correction term
(cid:104)v(cid:105)Dy =
1 + λ
(cid:16) 1
h(λζ − τ )
4πDR − hλζ−τ
8πD2
(cid:123)(cid:122)
(cid:17) .
(cid:19)
(cid:17)
(cid:125)
,
(11a)
(11b)
The approximation (cid:104)v(cid:105)0 is given by (cid:104)v(cid:105)2 without the loop
correction term. The derivation of Eq. (11) is outlined in
the Supplementary Material, in Section A5.
3These interactions are in the field-theoretic sense. In fact, the term for
(cid:96) = 0 already includes an arbitrary number of chemical interactions.
4
References
[1] Peter Bayley, Why microtubules grow and shrink, Na-
ture 363 (1993), 309.
[15] Jeffrey R. Kuhn and Thomas D. Pollard, Real-time
measurements of actin filament polymerization by to-
tal internal reflection fluorescence microscopy, Bio-
phys. J. 88 (2005), 1387 -- 1402.
[2] Otto G. Berg and Peter H. von Hippel, Diffusion-
controlled macromolecular interactions, Ann. Rev.
Biophys. Chem. 14 (1985), 131 -- 160.
[16] Tim Mitchison and Marc Kirschner, Dynamic insta-
bility of microtubule growth, Nature 312 (1984), 237 --
242.
[17] Richard M. Noyes, Effects of diffusion rates on chem-
ical kinetics, Prog. React. Kinet. 1 (1961), 129 -- 160.
[18] David J. Odde, Estimation of the diffusion-limited rate
of microtubule assembly, Biophys. J. 73 (1997), 88 --
95.
[19] L. Peliti, Path integral approach to birth-death pro-
cesses on a lattice, J. Phys.-Paris 46 (1985), 1469 --
1483.
[20] Thomas D. Pollard, Polymerization of adp-actin, J.
Cell Biol. 99 (1984), 769 -- 777.
[21] J. G. Skellam, The frequency distribution of the differ-
ence between two poisson variates belonging to differ-
ent populations, J. Roy. Stat. Soc. 109 (1946), 296.
[22] William Sutherland, A dynamical theory of diffusion
for non-electrolytes and the molecular mass of albium,
Phil. Mag. 9 (1905), 781 -- 785.
[23] Uwe C. Tauber, Critical dynamics, Cambridge Uni-
versity Press, 2014.
[24] M. v. Smoluchowski, Versuch einer mathematischen
theorie der koagulationskinetik kolloider losungen, Z.
Phys. Chem. 92 (1917), 129 -- 168.
[25] R.A. Walker, E.T. O'Brien, N.K. Pryer, M.F.
Soboeiro, W.A. Voter, H.P. Erickson, and E.D. Sla-
mon, Dynamic instability of individual microtubules
analyzed by video microscopy: Rate constants and
transition frequencies, J. Cell Biol. 107 (1988), 1437 --
1448.
[26] Michal Wieczorek, Sami Chaaban, and Gary J.
Brouhard, Macromolecular crowding pushes cat-
alyzed microtubule growth to near the theoretical
limit, Cell. Mol. Bioeng. 6 (2013), 383 -- 392.
[3] David Boal, Mechanics of the cell, Cambridge Univer-
sity Press, 2012.
[4] M. F. Carlier, R. Melki, D. Pantaloni, T.L. Hill,
and Y. Chen, Synchronous oscillations in microtubule
polymerization, Proc. Natl. Acad. Sci. Unit. States
Am. 84 (1987), 5257 -- 5261.
[5] Frank C. Collins, Diffusion in chemical reaction pro-
cesses and in the growth of colloid particles, J. Colloid
Sci. 5 (1950), 499 -- 505.
[6] Yi der Chen and Terrell Hill, Theoretical studies on
oscillations in microtubule polymerization, Proc. Natl.
Acad. Sci. Unit. States Am. 84 (1987), 8419 -- 8423.
[7] Masao Doi, Second quantization representation for
classical many-particle system, J. Phys. A: Math. Gen.
9 (1976), 1465 -- 1477.
[8] Detlev Drenkhahn and Thomas D. Pollard, Elonga-
tion of actin filaments is a diffusion-limited reaction
at the barbed end and is accelerated by inert macro-
molecules, J. Biol. Chem. 261 (1986), 12754 -- 12758.
[9] Carl Frieden, Actin and tubulin polymerization: The
use of kinetic methods to determine mechanism, Annu.
Rev. Biophys. Bio. 14 (1985), 189 -- 210.
[10] Ikuko Fujiwara, Shin Takahashi, Hisashi Tadakuma,
Takashi Funatsu, and Shin'ichi Ishiwata, Microscopic
analysis of polymerization dynamics with individual
actin filaments, Nat. Cell Biol. 4 (2002), 666 -- 673.
[11] Ikuko Fujiwara, Dimitrios Vavylonis, and Thomas D.
Pollard, Polymerization kinetics of adp- and adp-pi-
actin determined by fluorescence microscopy, Proc.
Natl. Acad. Sci. Unit. States Am. 104 (2007), 8827 --
8832.
[12] Melissa K. Gardner, Blake D. Charlebois, Imre M.
J´anosi, Jonathan Howard, Alan J. Hunt, and David J.
Odde, Rapid microtubule self-assembly kinetics, Cell
146 (2011), 582 -- 592.
[13] Frederick Gittes, Brian Mickey, Jilda Nettleton, and
Jonathan Howard, Flexural rigidity of microtubules
and actin filaments measured from thermal fluctua-
tions in shape, J. Cell Biol. 120 (1993), 923 -- 934.
[14] Kunkun Guo,
Julian Shillcock,
and Reinhard
Lipowsky, Self-assembly of actin monomers into long
filaments: Brownian dynamics simulations, J. Chem.
Phys. 131 (2009), 1 -- 11.
5
|
1305.5434 | 3 | 1305 | 2013-10-16T13:06:51 | Quantifying the sorting efficiency of self-propelled run-and-tumble swimmers by geometrical ratchets | [
"physics.bio-ph",
"cond-mat.soft"
] | Suitable asymmetric microstructures can be used to control the direction of motion in microorganism populations. This rectification process makes it possible to accumulate swimmers in a region of space or to sort different swimmers. Here we study numerically how the separation process depends on the specific motility strategies of the microorganisms involved. Crucial properties such as the separation efficiency and the separation time for two bacterial strains are precisely defined and evaluated. In particular, the sorting of two bacterial populations inoculated in a box consisting of a series of chambers separated by columns of asymmetric obstacles is investigated. We show how the sorting efficiency is enhanced by these obstacles and conclude that this kind of sorting can be efficiently used even when the involved populations differ only in one aspect of their swimming strategy. | physics.bio-ph | physics | Quantifying the sorting efficiency of self-propelled run-and-tumble swimmers by
geometrical ratchets
I. Berdakin1,2, A. V. Silhanek3, H. N. Moyano1, V. I. Marconi1,2, and C. A. Condat1,2
1Facultad de Matem´atica, Astronom´ıa y F´ısica,
Universidad Nacional de C´ordoba,
X5000HUA C´ordoba, Argentina.
2IFEG-CONICET,
X5000HUA C´ordoba, Argentina.
3D´epartement de Physique, Universit´e de Li`ege,
B-4000 Sart Tilman, Belgium.
Suitable asymmetric microstructures can be used to control the direction of motion in microor-
ganism populations. This rectification process makes it possible to accumulate swimmers in a region
of space or to sort different swimmers. Here we study numerically how the separation process de-
pends on the specific motility strategies of the microorganisms involved. Crucial properties such
as the separation efficiency and the separation time for two bacterial strains are precisely defined
and evaluated. In particular, the sorting of two bacterial populations inoculated in a box consisting
of a series of chambers separated by columns of asymmetric obstacles is investigated. We show
how the sorting efficiency is enhanced by these obstacles and conclude that this kind of sorting can
be efficiently used even when the involved populations differ only in one aspect of their swimming
strategy.
PACS numbers: 87.17.Jj, 87.17.Aa, 05.40.Fb
I.
INTRODUCTION
Self-propelled objects moving in confining environ-
ments at low Reynolds numbers exhibit interesting physi-
cal properties, some of which are not yet well understood
and deserve to be studied in view of their technologi-
cal applications. These objects range from artificial mi-
croswimmers that can be actuated upon by using applied
magnetic fields [1] to motile cancer [2, 3] and stem [4]
cells, to motile bacteria [5, 6] and spermatozoa [7, 8].
The study of their properties in confined regions has
been made possible by recent advances in microfabrica-
tion [9, 10] and observation [11, 12] techniques.
Aside from the intrinsic problems posed by the motion
of confined, self-propelled, run-and-tumble microorgan-
isms, there is strong interest in the biomedical and en-
gineering communities in efficiently controlling microor-
ganism motion. A physical, nondestructive method of
achieving this control is suggested by the geometrically-
induced guidance caused by the walls of asymmetric
openings, the funnels. This ratchet effect was first used
by Galajda and coworkers to generate an inhomogeneous
bacterial distribution [5]. The otherwise random motion
of bacteria was controlled, i.e rectified, by the funnels in
the box. Since bacteria swim parallel to the funnel walls,
it is easier for them to cross the barrier from the wide
to the narrow funnel opening than in the opposite direc-
tion. The funnels then define a preferred direction for
the swimmer motion, leading to an increase in the cell
concentration on one side of the box and a consequent
decrease on the other. This effect was also recently ob-
served in the puller eukaryote swimmer Chlamydomonas
reinhardtii
[13]. Interestingly, the rectification process
can be reversed if chemotactic or collective motions pre-
vail, as shown in Refs. [14, 15]. Various aspects of the
microorganism dynamics have been the subject of recent
studies [16 -- 23].
This rectification effect can be particularly useful when
there are mixtures of microorganisms exhibiting differ-
ent motility strategies.
In this connection it is worth
mentioning that various microfluidic techniques for sort-
ing motile microscopic objects have been developed in
the last few years. This is the case for the separation
of motile from non-motile sperm cells [24], the sorting
of E. coli by length [25], the use of self-driven artificial
microswimmers for the separation of binary mixtures of
colloids [26] and the study of the dynamics of several
kinds of particles combining asymmetric obstacles and a
time-dependent voltage [27]. One of their objectives is to
eliminate the cellular stress and damage associated with
alternative techniques such as centrifugation. Geometri-
cal sorting also avoids the use of applied fields or chemical
gradients, whose maintenance at scales of the millimeter
or longer is difficult [28]. Would it be also possible to
use the rectification effect to efficiently sort cells by their
swimming strategies? This is the question we would like
to answer in this paper.
Given their ubiquity, motile bacteria are of particular
interest. They swim by rotating thin helical filaments
called flagella; each flagellum is driven by a rotary mo-
tor powered by the flow of ions (H + or N a+) across
the cytoplasmic membrane. In order to take advantage
of chemotactic gradients, many of these bacteria have
evolved a run-and-tumble swimming strategy [29].
In
the case of the paradigmatic Escherichia coli, during the
run mode the flagella rotate counterclockwise and the mi-
3
1
0
2
t
c
O
6
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
3
v
4
3
4
5
.
5
0
3
1
:
v
i
X
r
a
croorganism moves in a forward, relatively straight direc-
tion, whereas during the tumble mode, one or more flag-
ella rotate clockwise and the bacterium is reoriented in a
new direction [30]. As shown forty years ago by Berg and
Brown for E. coli [31]: (A) runs are not strictly straight
due to rotational diffusion, (B) run lengths are exponen-
tially distributed, and (C) bacterial heading changes at
tumbles by less than 90◦, preserving some memory of the
previous run, a fact that is often neglected in theoretical
treatments. It is worth noting that some marine bacteria
show anti-persistency, in what is called a run-and-reverse
strategy [32, 33].
The case of bacterial directed motion under asym-
metrical geometrical confinement, first observed and ex-
plained in Ref. [5], was modeled phenomenologically in
Ref. [34]. The authors considered point-like swimming
bacteria following run-and-tumble dynamics with a con-
stant motor force magnitude and thermal fluctuations.
Although this model neglects the details of the swim-
mer dynamics, it reproduces the most important exper-
imental findings and has been an inspiration for fur-
ther theoretical work. In Ref. [35] the relation between
the ratchet effect and symmetry breaking by the fun-
nel array geometry was clarified.
It was shown there
that the break of time-reversal symmetry needed for rec-
tification is provided by the forced rotation of bacte-
ria when colliding with a wall, and not by the motor
force of bacteria. This is so because the break of time-
reversal symmetry provided by the bacterial motor is lost
at a coarse-grained level of diffusion where detailed bal-
ance is restored. Later, the influence of the specific dy-
namical properties, from (A) to (C), described by Berg
and Brown, on the accumulation of cells in presence of
asymmetric obstacles was studied in detail in Ref. [36].
This numerical analysis used experimental values of the
motility parameters. It was found that different swim-
ming strategies may yield very different microorganism
accumulation efficiencies, being measured as the device
capacity to concentrate cells (number of concentrated
cells/number of inoculated cells of the same type). We
summarize the main results of that work:
1. In unbounded environments there are two processes
that degrade the orientational correlation:
tum-
bling and rotational diffusion. The first is much
more important for systems with short runs, while
rotational diffusion gives the dominant contribu-
tion to memory loss in systems characterized by
long runs. These effects can be quantified by the
velocity correlation function.
2. A study of the mean square displacement in un-
bounded environments reveals that the transla-
tional diffusion coefficient DT decreases strongly as
the change-of-heading angle at a tumble increases.
Unless typical runs are very long, DT is approx-
imately given by its value in the absence of ro-
tational diffusion. These results indicate that, to
make accurate predictions about swimmer sorting,
2
TABLE I. Motility parameters of two different E. coli strains:
s1 corresponds to AW405 and s2 to CheC497 in Ref. [31].
Swimmer
¯v[µm/s] σv[µm/s] ¯φ[◦] σφ[◦] τ [s] DR[rad2/s]
s1 (wild type)
s2 (mutant)
14.2
20.0
3.4
4.9
68
33
36 0.86
15 6.30
0.18
0.06
it is necessary to consider the specific motility prop-
erties of the microorganisms involved.
3. When the dynamics of free swimmers is incor-
porated into a spatially constrained environment
(asymmetric geometry) long run lengths and small
tumble emergence angles lead to an increased cell
density near the walls and, consequently, to fast
net displacement in the easy ratchet direction. In
general, long permanence near the walls and suit-
able wall-of-funnels architecture, i.e., funnel walls
at least as long as the run length and funnel open-
ings of the order of the cell size, favor rectification
or cell concentration.
4. Increasing the average run duration, τ , increases
the time of permanence close to the walls, lead-
ing to enhanced rectification.
If τ is negligible,
the swimmers cannot take advantage of swimming
along the funnel walls and directed motion does not
ensue.
5. Increasing the average speed during the run and
decreasing the average change-of-heading angle at
a tumble, i.e. increasing persistence, enhance wall
accumulation and rectification.
6. Good agreement was obtained with available ex-
perimental data, specifically regarding the time of
rectification and the efficiency of a microfabricated
wall of funnel-shaped openings as the one used in
Ref. [5].
In this work, we use the improved phenomenological
model introduced by Berdakin et. al. in Ref. [36] to in-
vestigate the efficiency of asymmetric microarrays used as
sorting devices, and their dependence on the swimming
strategies of the microorganisms involved. Our objective
is to help to design good sorters, using a model that incor-
porates real motility parameters. In Section 2 we review
the computational model and define the quantities to be
calculated, such as the extraction time and the sorting
efficiency. In Section 3 we present our numerical results,
which are briefly discussed in the concluding section.
II. METHODS
The model. We study numerically a dilute system
of 2N0 microscopic self-propelled particles, the swim-
mers, moving under low Reynolds number conditions
and confined to a micro-patterned two-dimensional box
of size Lx×Ly. The box contains M identical, equidis-
tant columns of obstacles, each consisting of Nf openings
(asymmetric funnels), of gap size lg and wall length lf
(see Fig. 1(a)). We choose Nf = 3 and M from 5 to 20.
The M − 1 inner chambers and the last chamber have
all the same length lD = 150 µm. The length of the in-
oculation (leftmost) chamber is kept constant, lI = 450
µm, in order to have a fixed fractional occupied area to
define a high dilution initial condition at t = 0. The rele-
vant geometrical parameters are illustrated in Fig. 1: (a)
for the array configuration and (b) for the single-funnel
shape.
Each swimmer, whose location is given by a vector (cid:126)ri,
is represented by a soft disk of radius rs, moving in two
dimensions with speed vi and heading in the direction of
the unit vector v(Φi) = cos(Φi)i + sin(Φi)j.
In a con-
fined space, the swimmer dynamics is determined by the
overdamped equation of motion,
γ(cid:126)vi = (cid:126)F m
i + (cid:126)F sw
i + (cid:126)F s
i
(1)
where γ is the medium damping constant and the act-
ing forces are explained in detail below. The runs de-
scribed by Eq. (1) are interrupted by tumbles and af-
fected by rotational diffusion, all of which results in a
change of swimmer heading given by,
(cid:112)
∆Φi = ∆φχ + ν
2DR∆t(1 − χ)
(2)
where ν is a Gaussian-distributed random number, χ
is a state variable equal to 0 during a run and 1 dur-
ing a tumble, DR is the rotational diffusion coefficient
and ∆t the numerical integration time step. Aside from
the frictional damping, the forces operating on the mi-
croswimmer are:
1. Self propulsion. When starting to move under low
Reynolds number conditions, swimmers in an un-
bounded fluid get almost instantly to a constant
final speed.
In our model this constant speed is
given by F m/γ, where F m is the modulus of the
propelling force. The initial condition for the swim-
mer population speed is chosen from a normal dis-
tribution with mean ¯v and standard deviation σv.
Based on this distribution, each cell is assigned a
given speed at t = 0, which remains the same dur-
ing the whole simulation.
The heading of the swimmer is altered only by tum-
bling or rotational diffusion. Tumbles are assumed
to be instantaneous (real tumbles last 0.1 seconds
in mean about 10 % of the run time for wild type
E. coli and a smaller percentage for longer run bac-
teria - see Table I ). Tumbles result in a rotation,
∆φ, from the previous direction of motion, which
we consider Gaussian-distributed and centered at
¯φ with a width σφ, (see Table I and Ref. [36]).
3
Successive tumbles are spaced by almost straight
runs exponentially distributed with mean duration
τ . During a run, asymmetries in the self propul-
sion system and environmental fluctuations result
in deviations from a perfectly straight path. These
deviations are measured by the rotational diffusion
coefficient, DR, and included in our model via the
changes in the swimmer heading [38] expressed by
Eq. (2).
2. Interaction with the walls. It is modeled by a steric
repulsive force (cid:126)F sw
i
normal to the walls,
i = f sw(1 − rik/a)Θ(1 − rik/a)nk,
(cid:126)F sw
(3)
where f sw is the maximum strength of the force,
Θ is the step function, nk is a unit vector nor-
mal to the k-th wall, rik is the distance between
the i-th particle and the center of the k-th wall,
a = rs + w/2, and w is the wall width. Since
the swimming direction is unchanged during the
collision and the normal component of the propul-
sion force is counteracted by the repulsion of the
wall, the swimmer keeps lightly bouncing against
the wall. The component of (cid:126)F m that is parallel
to the wall propels the bacterium along the wall
with a reduced speed that is proportional to the
sine of the angle formed by the incidence direction
and nk. This phenomenological representation of
the interaction has the interesting property of re-
ducing the speed of the cells when they swim par-
allel to a wall without the need of adding an extra
parameter to the model. The speed reduction of
a bacterium swimming parallel to a wall has been
studied experimentally in Ref. [37]. Either a tum-
ble or rotational diffusion may allow the swimmer
to move away from the wall.
This interaction is responsible for the observed ac-
cumulation at the walls [38] and for the directed
motion and sorting of bacteria [5]. As remarked
with measurements of wall accumulation for bacte-
ria with different swimming strategies in Ref. [39],
the wild-type E. coli was significantly less attracted
to the surfaces than a mutant strain that does not
display tumbling.
3. A purely steric swimmer-swimmer repulsion of
maximum intensity f s,
ij = f s(1− (cid:126)rij /2rs)Θ(1− (cid:126)rik /2rs)(cid:126)rij,
(cid:126)F s
with (cid:126)rij = (cid:126)ri − (cid:126)rj. The hydrodynamic interac-
tion between microswimmers is not important at
very low swimmer concentrations [38], and we dis-
regard it here. Our approximations are buttressed
(4)
4
FIG. 1. Color online. (a) Sketch of the geometry of a 21−chamber microarray with M = 20 identical asymmetric funnel
columns and Nf = 3 (number of funnels). All chambers, except for the first inoculation chamber, lI , have the same length
lD = 150µm. (b) Single funnel geometry showing its relevant parameters:
lf , lg and θ. (c) Time evolution in minutes of
the normalized bacterial populations (Nch/N0) in the inoculation (leftmost) and last (rightmost) chambers where N0 is the
number of cells of each kind inoculated at t = 0 s. The bacterial populations are wild-type E. coli and a mutant used in [31]:
s1 (red/grey) and s2 (blue/black) respectively. Only s2 is present in the last chamber in the window of time considered, 26
min. The time to pick up s2 completely pure in ch21 is t∗ = 24.1 min and the s2 extraction efficiency is % = 85 %. (d) Time
evolution as in (c), but in the intermediate chambers. Chambers 17 to 19 are not shown but are all similar, without s1.
by recent measurements of cell-cell and cell-wall in-
teractions using E. coli, which show that thermal
and intrinsic stochasticity wash out the effects of
long-range fluid dynamics [40]. These experimen-
tal results imply that physical interactions between
bacteria are mainly determined by steric collisions
and lubrication forces.
A comparison with the model of reference [34] is in
order. In that reference the runs were assumed to have
a constant duration, all the swimmers moved with the
same speed and started in a completely random direc-
tion after each tumble (the emergence angles are uni-
formly distributed in [0, 2π]). Our model differs from
that of Ref. [34] in all these aspects. A further difference
is that we take into account rotational diffusion, which
was neglected in Ref. [34], where instead the center-of-
mass motion is affected by thermal random forces. These
changes were already introduced in Ref. [36] to obtain a
more faithful description of the observational facts.
Taking into account all the interactions described
above we arrive to our set of dynamical equations to
be solved numerically [36] for the 2N0 run-and-tumble
microswimmers. We assume that the mixed swimmers
population is initially randomly distributed in the inoc-
ulation chamber. Using a fourth order Runge-Kutta al-
gorithm we integrate the dynamical equations of motion
and we obtain the trajectories for each confined swim-
mer. The averages over realizations are later performed.
For simplicity we will always compare only two swimmer
strategies, a situation that is easiest to implement in the
laboratory using two different fluorescent markers. Of
è (a)(b)(c )(d)course more than two swimmers could also be sorted as
shown in Ref. [36]. Table I specifies the motility param-
eters of the swimmers simulated in this work, the wild
type E. coli, s1, and a faster, less frequently tumbling
mutant, s2. The radius of the soft disks is taken to be
rs = 0.5 µm for all swimmers. The optimal single-funnel
geometric parameters for an efficient rectification of wild-
type E. coli, s1, were found in Ref. [36] to be lg = 2 µm,
lf = 30 µm, and θ = 68◦, so we keep these parame-
ters fixed for all simulations. The box width used here is
Ly = 80 µm and the wall width is taken to be w = 2 µm
for both, the box walls and the funnel walls. The width of
the inoculation chamber, lI = 450 µm, has been chosen
to keep an initially low swimmer density. The number of
swimmers, N0, of each strain is adapted to maintain an
initial occupied area fraction of 0.05 at the inoculation
chamber for all array geometries. If γ = 6πηrs, the fric-
tional drag coefficient, rs = 0.5 µm and η = 10−2 poise
(the viscosity of water at 20◦C), then γ ∼ 9.425 x 10−6
g/s. Under these conditions, the strength of the motor
force of a bacterium swimming at 20 µm/s is 0.17 pN .
The magnitudes of the forces acting upon the swimmers
are f s = 200 and f sw = 300 in units of γ, equivalent,
respectively, to ten and fifteen times the force exerted by
the motor on the fluid at 20 µm/s. With this choice in
our phenomenological model bacteria penetrate no more
than 10 % of rs inside walls or other bacteria.
Calculated quantities. With the aim of quantifying
the efficiency of the sorters we propose two parameters as
convenient indicators of the separation process: (a) the
separation time, t∗, defined as the time elapsed between
the arrival of the first swimmer in the fast, s2, class, and
that of the first swimmer in the slow, s1, class, to the
last chamber (chamber from where a pure cell population
could be extracted or concentrated), and (b) the separa-
tion efficiency, %, which we define as the fraction of the
fastest type that has arrived at the last chamber by the
time t∗. It is convenient to define the percent extraction
efficiency as follows,
% = 100
NF (t∗)
N0
,
(5)
being NF (t∗) being the number of swimmers of the fast
species that is present in the last chamber at t∗, when its
purity is still 100%.
III. RESULTS
We first consider a sample with M = 20 funnel
columns and two homogeneous bacterial distributions ini-
tially inoculated in the first chamber. These bacteria
are wild-type E. coli and a mutant studied by Berg and
Brown [31]; their characteristic dynamical parameters are
specified in Table I. We compute the variations of the to-
tal bacterial populations of each mutant in the first and
the last chambers, which are shown in Fig. 1(c). After
30 independent realizations, the average separation time
5
for this system is t∗ = 16 min and the average sepa-
ration efficiency for the mutant s2 is % = 67 ± 19%.
Three factors contribute to the high extraction efficiency
for this mutant: its higher average speed, its higher per-
low ¯φ, and, mainly, the longer duration of
sistence, i.e.
the runs, which increases both DT and the contact time
with the rectifying walls. The advance of both popula-
tions through the various chambers is shown in Fig. 1(d),
where we see that the purification process improves with
successive chambers. From chamber 12 onwards, we also
observe that the time evolution of the s2 pulse (blue)
is almost position-independent until it reaches the last
column.
Instantaneous snapshots of the bacterial pop-
ulations considered in Fig. 1 are shown, as functions of
time, in Fig. 2(a), where they are seen to start from a
uniform distribution in the inoculation chamber and ad-
vance at different rates in the easy ratchet direction. A
comparison between corresponding panels in Figs. 2(b)
and 2(c) shows clearly how these rates are enhanced by
the ratchet geometry of the column array, giving an es-
timated 5 µm/s drift velocity for the s2 population, five
times larger than that found for s1. As a result of these
different velocities inside the box, both populations are
soon largely separated and can be readily sorted out.
It is interesting to compare what happens in the spe-
cially designed box, an array of funnel columns, with the
result obtained in a single channel with the same area
and clean of obstacles, when the bacterial populations
are subject to the same initial conditions. In the clean
box, as the histograms in Fig. 2(b) show, the fast type
also moves forward first, in part taking advantage of its
longer runs along the side walls, but the separation is
much less efficient than for the funnel-containing box,
for which at t = 20 min there is no s1 swimmer from
chamber 16 to 21. Purification is complete there. For
the particular realization represented in Fig. 2(b) the ex-
traction time and extraction efficiency are, respectively,
t∗ = 24.1 min and % = 85%. At very long times, a
uniform distribution is expected for the single-chamber
configuration, while an exponentially increasing popula-
tion of each bacterial type is expected in the specially
designed box. This exponential increase is responsible
for the high concentrations near the end of the array of
columns, which permit the extraction of a high fraction
of the first bacterial type arriving there. This situation is
clearly shown in Fig. 2(c), where it is also possible to ob-
serve s2 concentration spikes where the obstacle columns
are located. The stronger tendency of s2 to concentrate
near the walls, as compared to s1, was recently studied
in detail [36].
Now we study the sorter efficiency of two swimmers
(one real and the second real or artificial) as a function
of the specific dynamical parameters characterizing the
microswimmers. In Fig. 3 we show the extraction times
and sorting efficiencies for two swimmers, one of which is
wild-type E. coli, s1, and the other sx, for which a sin-
gle motility parameter is changed. From Figs. 3(a) and
3(b), for which only the mutant speed was changed, we
6
FIG. 3. Color online. (a) Extraction or pick up time and
(b) sorting efficiency for the fastest swimmer to reach the last
chamber when we simultaneously simulate wild-type E. coli,
s1, and a mutant, sx, for which only the average run speed
is changed. (c) and (d): the same quantities when only the
average tumbling angle ¯φx of the mutant is changed. Inset:
sorting efficiency when only the mean run duration, τx, of
the second swimmer is changed. Note the different vertical
scales between (b) and (d). Here we use a smaller array with
M = 10.
changing the mean speed, the corresponding sorting ef-
ficiencies, panel (d), are markedly lower.
In this case,
we compared a hypothetical swimmer sx, whose average
tumbling angle ¯φx is modified, with the wild type, for
which ¯φ1 = 68◦. Easiest to separate are the persistent-
walk bacteria, for which ¯φx = 0◦ and the run-and-reverse
bacteria, for which ¯φx = 180◦. It is worth noting that
large t∗ does not necessarily mean large %. For example,
if ¯φx = 180◦, t∗ = 16 min, but % is only 5%, a relatively
low value when compared with % = 20% that results for
¯φx = 0◦, for which t∗ is only 7 min.
IV. DISCUSSION
We have investigated arrays of asymmetric-funnel
columns built for the purpose of concentrating or sort-
ing out one type of self-propelled swimmer in a run-and-
tumble microorganism mixture. As characteristic param-
eters to measure the suitability of a given architecture, we
introduced the extraction time and the sorting efficiency.
The first is important because it gives us the length of
the temporal window available to the experimentalist to
pick up the chosen strain, but does not tell us anything
about the number of swimmers ready to be extracted.
This is given by the separation or sorting efficiency.
The separation efficiency depends both on the motil-
ity parameters of the swimmers and on the geometrical
dimensions of the device, which we can modify accord-
ing to the swimmer types we are dealing with. Here we
have considered the competition between swimmers hav-
FIG. 2. Color online. (a) Snapshots illustrating the separa-
tion of the two bacterial types considered in Fig. 1. After
being uniformly inoculated in the first chamber, both popu-
lations are rapidly segregated. (b) Comparative profiles for
the spatial evolution of the s1 (red/grey) and s2 (blue/black)
bacterial populations for a clean box (upper histograms) and
for the 21-chamber box of the same overall size (lower his-
tograms) at the indicated times. The 21-chamber box is a far
more efficient extraction device than the single-chamber box.
see that, when vx < v1 = 14.2 µm/s (shadowed region),
the wild-type bacteria arrive first and can be purified dur-
ing a time t∗. This purification window is, for instance,
of 20.3 min if vx = 8 µm/s. The window gets narrower
when vx is close to v1, but grows monotonically when
vx > v1. Similarly, the sorting efficiency has a minimum
when vx = v1 but increases with the difference between
bacterial speeds. We can purify 18% of 30 µm/s mutants
and we have 7.5 minutes to do it. This behavior was to
be expected, since a faster bacterium diffuses farther, and
more importantly, can take advantage of longer runs par-
allel to the walls. The saturation value of t∗ is given by
the average time it would take the "slow" bacterial strain
to travel from the inoculation chamber to the end of the
box (this would be the separation time for a hypothet-
ical infinitely fast strain). The inset in panel (b) shows
that varying the mean run duration has an effect qual-
itatively similar to changing the mean speed (τ1 = 0.86
s). Although changing the mean tumbling angle, see
panel (c), yields extraction times of the same order as
205080110counts(b)t=1min2050801100.20.50.8counts(c)t=6min0.20.50.8t=10min0.20.50.8t=20min0.20.50.8x/Lx 4 8 12 16 20t* [min](a) 0 20 40 60 801030507090ε% -vsx [µm/s](b)0.010.1110110-τsx [s]ε% 4 8 12 16 20(c)3070110150 0 5 10 15 20-φsx [degrees](d)ing different intrinsic dynamical properties. Currently,
we are working out in detail the effect of modifications in
the geometrical array parameters that define the asym-
metric confining system.
The following are some predictions from our study:
• Asymmetric funnel arrays are capable of sorting di-
luted distributions of run-and-tumble swimmers in
a controlled way, enhancing the efficiency obtained
using a box free of geometrical constraints.
• A sizable fraction of the chosen swimmers can
be 100% purified even if the original mixture is
composed of swimmers that are dynamically only
slightly different.
• In general, unless the motility properties of the
swimmers are very similar, for M of the order of
10 the extraction time should be long enough to
allow the experimentalist to purify the sample.
In our simulations we did not include fluid flow, so that
the net bacterial motion from left to right is solely due to
funnel asymmetry. Under flow our results would be very
different. Flow in a narrow channel is known to change
the accumulation of cells on the walls and even to cause
upstream swimming. Moreover, the response to flow de-
pends upon the tumbling rate of the cells [12, 41]. We
could hypothesize that flow may lower the sorting effi-
ciency when pointing in the easy ratchet direction (left-
right) and enhance the efficiency otherwise, but this is
something that deserves careful study.
In this paper our goal was to efficiently sort swimmers
at low concentrations as experiments in view of how tech-
nological applications in this field are generally made.
But what would happen at high concentrations? One
way to look at these problems is to adapt the well-known
Vicsek's model [42 -- 45]. Hydrodynamic equations have
also been obtained in the high density limit using a Boltz-
mann approach [45] and through the coarse-graining of
the microscopic dynamics [46]. Recently, Drocco and
coworkers [15] added steric repulsion to the Vicsek flock-
ing algorithm and studied the motion of self-propelled
particles in a confining microenvironment such as the one
considered in this paper. These authors found rectifica-
tion effects induced by the high particle concentration in
7
the absence of preferential motion along the walls. The
nature of this rectification process is therefore quite dif-
ferent from the one we have considered here and opens
the way to the analysis of a possibly rich phenomenology
and other types of applications.
Another extension of the studies in this paper that
would be specially profitable in the case of the smallest
self-propelling microorganisms could be made by explic-
itly considering the influence of passive and active fluctu-
ations on the system behavior. The impact of the differ-
ent fluctuation types on the collective dynamics of active
Brownian particles with velocity alignment has already
been studied [47, 48]. Our understanding of microswim-
mer dynamics would be enhanced by the analysis of the
behavior of these active particle systems in asymmetric
confining microarchitectures.
To summarize, the purpose of this paper was twofold:
First, to introduce new definitions, those of extraction ef-
ficiency and of separation time, which are advantageous
to quantify how effective is a given microarchitecture to
sort different types of run-and-tumble self-propelling mi-
croorganisms. Second, to specifically investigate how
these microorganisms can be sorted by their motility
strategies. We have shown how testable predictions can
be made using realistic bacterial parameter values. These
predictions can be very useful to design efficient microflu-
idic devices. We further point out that, although run-
and-tumble strategies are common in the bacterial world,
this type of motion is not restricted to bacteria. The lo-
comotion of the unicellular alga chlamydomonas exhibits,
in the dark, nearly straight swimming runs interrupted
by abrupt changes in direction. The run distributions are
exponentially distributed, with ¯τ = 11.2 s [49]. Conse-
quently, the dynamics of this eukaryote are likely to be
describable by the model discussed in this paper as well.
Our numerical calculations can be easily generalized to
include the possibility of bacterial birth/death during the
experiment, work we have in progress.
Acknowledgments
This work was supported by CONICET (PIP 112-
200801-00772 and PIP 112-201101-00213) and SeCyT-
UNC (Projects No. 05/B354 and No. 30720110101526)
(Argentina),
(Belgium)-CONICET, FWO-
MINCyT FW/09/04 and KU Leuven-UNC bilateral
projects.
FNRS
[1] E. E. Keaveny, S. W. Walker, M. J. Shelley, Nano Lett.
[7] J. Elgeti, U. B. Kaupp, G. Gompper, Biophys. J. 99,
13, 531 (2013)
1018 (2010)
[2] G. Mahmud et al., Nature Phys. 5, 606 (2009)
[3] K. Konstantopoulus, P. Wu, D. Wirtz, Biophys. J. 104,
279 (2013)
[4] R. Peng, X. Yao, J. Ding, Biomaterials. 32, 8048 (2011)
[5] P. Galajda, J. Keymer, P. Chaikin, R. Austin, J.
[8] P. Denissenko, V. Kantsler, D. J. Smith, J. Kirkman-
Brown, Proc. Natl. Acad. Sci. USA 109, 8007 (2012)
[9] T. M. Squires, S. R. Quake, Rev. Mod. Phys. 77, 977
(2005)
[10] K. Leung et al., Proc. Natl. Acad. Sci. USA 109, 20
Bacteriol. 189, 8704 (2007)
(2012)
[6] S. Y. Kim, E. S. Lee, H. J. Lee, S. Y. Lee, S. K. Lee,
T. Kim, J. Micromech. Microeng. 20, 0950061 (2010)
[11] L. G. Wilson et al., Phys. Rev. Lett. 106, 018101 (2011)
[12] E. Altshuler, G. Mino, C. P´erez-Penichet, L. del R´ıo,
8
A. Lindner, A. Rousselet, E. Cl´ement, Soft Matter 9,
1864 (2013)
[31] H. C. Berg, D. A. Brown, Nature 239, 500 (1972)
[32] G.M. Barbara, J.G. Mitchell, FEMS Microbiol. Ecol. 44,
[13] V. Kantsler, J. Dunkel, M. Polin, R. E. Goldstein, Proc.
79 (2003)
Natl. Acad. Sci. USA 110, 4 (2013)
[33] L. Xie, T. Altindal, S. Chattopadhyay, X.L. Wu, Proc.
[14] G. Lambert, D. Liao, R. H. Austin, Phys. Rev. Lett. 104,
Natl. Acad. Sci. USA 108, 22462251 (2011)
168102 (2010)
[34] M. Wan, C. J. O. Reichhardt, Z. Nussinov, C. Reich-
[15] J. A. Drocco, C. J. O. Reichhardt, C. Reichhardt, Phys.
hardt, Phys. Rev. Lett. 101, 018102 (2008)
Rev. E 85, 056102 (2012)
[16] C. A. Condat, J. Jackle, S. A. Mench´on. Phys. Rev.
E 72, 021909, (2005)
[17] F. Peruani, L. G. Morelli, Phys. Rev. Lett. 99, 010602
(2007)
[35] J. Tailleur, M. E. Cates, Eur. Phys. Lett. 86, 60002 (2009)
[36] I. Berdakin,Y. Jeyaram, V. V. Moshchalkov, L. Venken,
S. Dierckx, S. J. Vanderleyden, A. V. Silhanek,
C. A. Condat, V. I. Marconi, Phys. Rev. E 87, 052702
(2013)
[18] V. Garcia, M. Birbaumer, F Schweitzer, Eur. Phys. J. B
[37] P. D. Frymier, R. M. Ford, H. C. Berg, P. T. Cummingssi,
82, 235 (2011)
[19] M. E. Di Salvo, C. A. Condat, Phys. Rev. E 86, 061907
(2012)
[20] G. Gr´egoire, H. Chat´e, Phys. Rev. Lett. 92, 025702
(2004)
[21] E. Bertin, M. Droz, G. Gr´egoire, J. Phys. A 42, 445001
(2009)
[22] C. A. Weber et al., Phys. Rev. Lett. 110, 208001 (2013)
[23] P. Romanczuk, L. Schimansky-Geier, Phys. Rev. Lett.
106, 230601 (2011)
[24] B. S. Cho, T. G. Schuster, X. M. Zhu, D. Chang,
G. D. Smith, S. Takayama, Anal. Chem. 75, 4671 (2003)
[25] S. E. Hulme, W. R. DiLuzio, S. S. Shevkoplyas,
L. Turner, M. Mayer, H. C. Berg, G. M. Whitesides,
Lab on a Chip 8, 1888 (2008)
Proc. Natl. Acad. Sci. USA 92,6195-9 (1995)
[38] G. Li, J. X. Tang, Phys. Rev. Lett. 103, 078101 (2009)
[39] G. Mino, T. E. Mallouk, T. Darnige, M. Hoyos,
J. Dauchet, J. Dunstan, R. Soto, Y. Wang, A. Rous-
selet, E. Cl´ement, Phys. Rev. Lett. 106, 0481021 (2011)
[40] K. Drescher, J. Dunkel, L. H. Cisneros, S. Ganguly,
R. E. Goldstein, Proc. Natl. Acad. Sci. USA 108,10940
(2011)
[41] A. Costanzo, R. Di Leonardo, G. Ruocco, L. Angelani,
J. Phys.: Condens. Matter 24 065101 (2012)
[42] T. Vicsek, A. Czirok, E. Ben-Jacob,
O. Shochet, Phys. Rev. Lett. 75, 1226 (1995)
I. Cohen,
[43] G. Gr´egoire, H. Chat´e, Y. Tu, Physica D 181, 157 (2003).
[44] H. Chat´e, F. Ginelli ,G. Gr´egoire, F. Raynaud, Phys.
Rev. E 77, 046113 (2008)
[26] W. Yang, V. R. Misko, K. Nelissen, M. Kong,
[45] E. Bertin, M. Droz, G. Gr´egoire, Phys. Rev. E 74, 022101
F. M. Peeters, Soft Matter 8, 5175 (2012)
(2006)
[27] L. Bogunovic, R. Eichhorn, J. Regtmeier, D. Ansel-
metti, P. Reimann, Soft Matter 8, 3900 (2012)
[46] T. Ihle, Phys. Rev. E 83, 030901(R) (2011).
[47] R. Grossmann, L. Schimansky-Geier, P. Romanczuk,
[28] J. Voldman, Nature Phys. 5, 536 (2009)
[29] H. C. Berg, Random Walks in Biology (Princeton Uni-
versity Press, Princeton, New Jersey, 1993)
[30] H. C. Berg, E. coli in motion (Springer, New York,
2004)
New Journal of Physics14, 073033 (2012)
[48] P. Romanczuk, M. Baer, W. Ebeling, B. Lindner,
L. Schimansky-Geier, Eur. Phys. J. Special Topics 202,
1 (2012)
[49] M. Polin, I. Tuval, K. Drescher, J. P. Gollub,
R. E. Goldstein, Science 325, 487 (2009)
|
1203.3005 | 1 | 1203 | 2012-03-14T06:11:59 | On the Theory of Dielectric Spectroscopy of Protein Solutions | [
"physics.bio-ph",
"cond-mat.soft"
] | We present a theory of the dielectric response of a solution containing large solutes, of a nanometer size, in a molecular solvent. It combines the molecular dipole moment of the solute with the polarization of a large subensemble of solvent molecules at the solute-solvent interface. The goal of the theory is two-fold: (i) to formulate the problem of the dielectric response avoiding the reliance on the cavity-field concepts of dielectric theories and (ii) to separate the non-additive polarization of the interface, jointly produced by the external field of the laboratory experiment and the solute, from specific solute-solvent interactions contributing to the dielectric signal. The theory is applied to experimentally reported frequency-dependent dielectric spectra of lysozyme in solution. The analysis of the data in the broad range of frequencies up to 700 GHz shows that the cavity field susceptibility, critical for the theory formulation, is consistent with the prediction of Maxwell's electrostatics in the frequency range of 10--200 GHz, but deviates from it outside this range. In particular, it becomes much smaller then the Maxwell result and shifts to negative values at small frequencies. The latter observation implies a dia-electric response, or negative dielectrophoresis, of hydrated lysozyme. It also implies that the effective protein dipole recorded by dielectric spectroscopy is much smaller than the value calculated from protein's charge distribution. We suggest an empirical equation that describes both the increment of the static dielectric constat and the decrement of the Debye water peak with increasing protein concentration. It gives fair agreement with broad-band dispersion and loss spectra of protein solutions, but misses the delta-dispersion region. | physics.bio-ph | physics |
On the Theory of Dielectric Spectroscopy of Protein
Solutions
Dmitry V. Matyushov
Center for Biological Physics, Arizona State University, PO Box 871504, Tempe, AZ
85287-1504
E-mail: [email protected]
Abstract. We present a theory of the dielectric response of a solution containing
large solutes, of a nanometer size, in a molecular solvent. It combines the molecular
dipole moment of the solute with the polarization of a large subensemble of solvent
molecules at the solute-solvent interface. The goal of the theory is two-fold: (i) to
formulate the problem of the dielectric response avoiding the reliance on the cavity-
field concepts of dielectric theories and (ii) to separate the non-additive polarization
of the interface, jointly produced by the external field of the laboratory experiment
and the solute, from specific solute-solvent interactions contributing to the dielectric
signal. The theory is applied to experimentally reported frequency-dependent dielectric
spectra of lysozyme in solution. The analysis of the data in the broad range of
frequencies up to 700 GHz shows that the cavity field susceptibility, critical for the
theory formulation, is consistent with the prediction of Maxwell's electrostatics in the
frequency range of 10 -- 200 GHz, but deviates from it outside this range. In particular,
it becomes much smaller then the Maxwell result and shifts to negative values at
small frequencies. The latter observation implies a dia-electric response, or negative
dielectrophoresis, of hydrated lysozyme. It also implies that the effective protein dipole
recorded by dielectric spectroscopy is much smaller than the value calculated from
protein's charge distribution. We suggest an empirical equation that describes both
the increment of the static dielectric constat and the decrement of the Debye water
peak with increasing protein concentration. It gives fair agreement with broad-band
dispersion and loss spectra of protein solutions, but misses the δ-dispersion region.
Keywords: protein electrostatics; dielectric spectroscopy; solution; nanoscale interface
PACS numbers: 87.15.-v, 87.15.He, 87.15.By, 87.10.Pq
Submitted to: J. Phys.: Condens. Matter
Protein dielectrics
1. Introduction
2
Dielectric spectroscopy is a linear response technique, monitoring the dynamics of the
dipole moment of a macroscopic sample of a polarizable material [1, 2]. While it is highly
sensitive and provides wealth of information about the dynamics of polarization modes
active in a medium, the interpretation and the assignment of the observed relaxation
processes often require theoretical approaches.
The standard theoretical tool to study mixtures is the Maxwell-Wagner theory [2, 3]
and its modifications, also in terms of effective-medium approaches [4]. All these theories
assume that macroscopic dielectric constants can be assigned to all components of the
mixture. This often becomes a significant oversimplification when highly heterogeneous
solutes of nanometer dimension, such as hydrated proteins, are involved [5, 6]. The
description of the polar response in terms of the molecular charge distribution is more
accurate for these solutes [7]. Given the length-scale of the external field variation in
a typical dielectric or light-absorption experiment, the overall charge and the dipole
moment are the two main multipoles to consider [8].
Once a dipole is assigned to a protein, one might assume that standard models of
dipolar liquids [9, 10], involving the statistics and dynamics of molecular dipoles, can
be directly extended to study protein solutions. One has, however, to recognize that
proteins, and other solutes of similar dimension, possess an extended interface with a
molecular solvent, such as water, which is absent in the case of mixtures composed
of molecules of comparable size. The interface of a hydrated protein involves a large
number, ∼ 300 − 500, water molecules only in the first hydration layer. Given that
the perturbation of the water polarization propagates at least into the second hydration
layer [11, 12], the actual size of the protein-water interface is significantly larger.
These new physical realities pose the requirement to develop new theoretical
approaches to describe the polar response of protein solutions. The key question for
this development is how to extend the classical theories of polar response of molecular
dipoles into the realm of large solutes with an extended interfaces. The key parameter
for the development of dielectric theories is the Onsager cavity (or directing) field [13]
producing the torque on the molecular dipole when the macroscopic sample is placed in
an external electric field [14]. The standard result of the classical theories is that the
field of external charges E0 is screened by the polarization of the interface to the cavity
field [14, 15]
Ec =
3
2ǫs + 1
E0,
(1)
where ǫs is the dielectric constant of the solvent. This cavity field then directly leads
to the Onsager mean-field [16] equation for the dielectric constant and, when mutual
short-range correlations of dipoles are included, to the Onsager-Kirkwood relation [14].
The problem one faces in an attempt to describe a mixture of nanometer-size solutes
with a molecular solvent is that there is no analog of either of these two equations.
The fundamental line of inquiry here is whether one can extend equation (1), or its
Protein dielectrics
3
analog, to such mixtures, or a new set of rules is required. This question is nontrivial to
answer, even though some initial computer simulations indicate that, indeed, polarized
nanoscale interfaces follow rules different from those established for cavities carved in
dielectrics [17, 18, 12]. The simulations are however limited by the nanosecond range of
time-scales. The question of what is the polar response of a nanoscale interface at low
frequencies remains therefore open.
This study aims to address this question by analyzing recent dielectric data obtained
for solutions of lysozyme in water [19, 20]. We first develop a general formalism that
does not anticipate any particular solution for the local field acting on the protein dipole.
As a result of this analysis, we arrive at a surprising conclusion that the hydration layers
of the protein screen its dipole even more substantially than anticipated by the standard
result for a dielectric cavity given by equation (1).
We start with introducing the polarization of the solute-solvent interface by the
combined effect of the external electric field and the solute dipole moments. This
interfacial polarization integrates into an interface dipole moment, which is assigned to
each solute even in the absence of its own dipole. This development leads to the equation
for the dielectric constant of an ideal solution of dielectric voids inside the polar liquid.
We show that this equation is quite useful in describing the high-frequency dielectric
response of a real solution, when the relaxation of the solute dipoles is dynamically
frozen. We then proceed to a mixture of polar solutes in a polar liquid. Here, the
cross-correlations of the solute and solvent dipoles [21] are expressed in terms of the
cavity-field susceptibility, which can take different forms depending on the microscopic
structure of the water layer interfacing the solute [18].
2. Dielectric Response of a Mixture
Dissolving a polar solute in a polar solvent leads to two distinct effects on the response
of the medium to an external electric field. The first effect is the exclusion of the
solvent from the volume of the solute. The second effect is the response of the charge
distribution within solute to the orienting torque of the external field. The two effects are
entangled in the polarization of the interface by the solute charges and by the external
field. However, their contributions to the overall dielectric response of the solution can
be separated in the frequency domain. Since they also originate from distinct physical
interactions, repulsive expulsion on the one hand and electrostatic interactions on the
other, we start with considering the effect of the solute excluded volume and then add
the contribution of the solute dipole moment to the dielectric response of the solution.
2.1. Non-polar Solutes in a Polar Solvent
Excluding the solvent from the solute volume creates the solute-solvent interface.
From the standard viewpoint of dielectric theories, any interface carries an interfacial
polarization when the solution is placed in a uniform field of external charges (capacitor
Protein dielectrics
4
plates of the dielectric experiment). The polarization of the interface is described in
the Maxwell theory of dielectrics by the surface charge density [15]. It is given as the
projection of the dipolar polarization of the dielectric PS at the dividing surface S on the
outward normal to the surface nS. The surface charge density then becomes σP = n·PS.
This charge density integrates to a dipole moment of the interface
Mint
0 = ZS
rSσP (rS)dS,
(2)
where the surface integral is taken over the closed surface S enveloping the solute.
The interface dipole polarizes the surrounding dielectric by its own electric field such
that the inhomogeneous Maxwell field E(r) around the solute is a sum of the uniform
Maxwell field of the external charges ǫ−1
s E0 and the dipolar field of the polarized interface
E(r) = ǫ−1
s E0 +Xj
T(r − rj) · Mint
0 .
(3)
Here, T(r − rj) is the dipolar tensor describing the electric field at point r inside the
solvent by a point dipole placed at rj; the sum runs over N0 solutes with coordinates
rj.
The Maxwell field E(r) polarizes the liquid, with the resulting local inhomogeneous
polarization P(r) = (4π)−1(ǫs − 1)E(r), decaying to the homogeneous polarization P of
the external charges far from the solute-solvent interface. The overall dipole created in
the solution is the integral of P(r) over the volume Ω occupied by the solvent
Mmix = ZΩ
P(r)dr.
(4)
Here, subscript "mix" identifies the solvent-solute mixture. Assuming that the
interfacial dipoles of solutes are independent of each other, one gets [12]
Mmix = Mhom − N0Ω0P − (2/3)(ǫs − 1)N0Mint
0 .
(5)
Here, Mhom = V P is the dipole moment of the corresponding homogeneous (without
solutes) polarized solvent and Ω0 is the volume of the solute; P = (4π)−1(1 − ǫ−1
s )E0
is the polarization of the homogeneous solvent. The second summand in equation (5)
represents the dipole moment cut from the liquid by inserting N0 voids. Finally, the
last term is an additional polarization induced in the surrounding liquid by the surface
charge density σP .
The value of the interface solute dipole M int
0 will depend on the specifics of the
solute-solvent interactions and the local polarization of the solvent created by these
interactions. While it is a complex function of the entire mosaic of pairwise solute-
solvent interactions for a realistic solute, an estimate of this parameter can be obtained
from dielectric theories for a spherical void in a dielectric. The interface dipole reads in
this case [15]
MM
0 = −3Ω0P/(2ǫs + 1),
(6)
Protein dielectrics
5
where the subscript "M" specifies Maxwell's electrostatics of a dividing surface not
affected by local solute-solvent interactions. In order to quantify deviations from this
generic result, one can introduce the ratio
α = M int
0 /M M
0 .
(7)
The dipole moment of the mixture is related to the mixture dielectric constant ǫmix
mix)E0. One then obtains for the dielectric constant of the
as Mmix/V = (4π)−1(1 − ǫ−1
mixture
ǫs
ǫmix
= 1 + η0(ǫs − 1)(cid:20)1 − 2α
ǫs − 1
2ǫs + 1(cid:21) + R1(η0),
(8)
where η0 = N0Ω0/V is the volume fraction of the solutes in the sample with the overall
volume V . We have put an extra term R1(η0) in the above equation to indicate terms
non-linear in the volume fraction that appear in the dielectric constant when mutual
polarization of the interfacial dipoles is taken into account [22]. Similar non-linear terms
appear in the response of a mixture of water with dipolar solutes discussed below. There
is presently no consistent formalism to include these effects and we neglect them at the
current stage of the theory development recognizing that the theory might run into
conflict with the data collected for concentrated solutions.
If the Maxwell result for a void in a dielectric holds, α = 1 and the dielectric
constant of the mixture becomes
ǫs
ǫmix
= 1 + η0
3(ǫs − 1)
2ǫs + 1
.
(9)
Equation (8), with R1(η0) omitted, and (9) describe the dielectric constant of an ideal
mixture of non-polar solutes and a polar solvent. Equation (9) also reduces to the
standard result of the Maxwell-Wagner theory in the limit of low volume fraction of the
solutes [2, 3]. One can also account for the electronic polarizability of the protein not
mentioned so far. If the refractive index np can be assigned to the protein, one needs
only to realize that the boundary conditions of the dielectric theories are sensitive to the
ratio of the two dielectric constants at the dividing surface, ǫs/n2
p. Equation (9) then
extends to
ǫs
ǫmix
= 1 + η0
3(ǫs − n2
p)
2ǫs + n2
p
.
(10)
Equation (8) can be alternatively written in terms of the cavity field Ec inside a
spherical void in a uniformly polarized liquid. The electric field inside the cavity is
proportional to the external field, with the susceptibility χc = Ec/E0. In terms of this
susceptibility, (8) becomes [18]
ǫs
ǫmix
= 1 + 3η0 [χcǫs − 1] .
The standard prescription of Maxwell's theory of dielectrics predicts [14, 23]
χM
c =
3
2ǫs + 1
.
(11)
(12)
Protein dielectrics
6
The connection between the susceptibility χc and the parameter α (equation (7))
that is required to obtain equation (11) from equation (8) is derived from the following
arguments. The polarization P(r) in the solvent, induced by the Maxwell field given by
equation (3), creates a non-vanishing electric field inside the solute that is given by the
equation
Ec = E0 +ZΩ
T(r) · P(r)dr.
(13)
Upon substitution of equation (3) into this relation, one arrives at the connection
between χc and α
3ǫsχc = ǫs + 2 − α
2(ǫs − 1)
2ǫs + 1
.
(14)
Combining equations (8) and (14), one arrives at equation (11).
2.2. Polar Solutes in a Polar Solvent
When a solute carries dipole moment m0, it aligns along the external field such that the
average dipole hm0iE in a weak external field is given by linear susceptibility [23] χ0
hm0iE = χ0Ω0E0,
where h. . .iE denotes an ensemble average in the presence of the external field and
χ0 = χ00 + χ0s = (β/3Ω0)hδm0 · δMmixi.
(15)
(16)
In this equation, δm0 = m0 − hm0i and δMmix = Mmix − hMmixi are the deviations
of the solute dipole and the dipole of the sample Mmix from their average values and
β = 1/(kBT ) is the inverse temperature.
The solute susceptibility in equation (16) is split into the self, χ00, and solute-
solvent, χ0s, parts. The former is given by the variance of a single solute dipole
χ00 = (β/3Ω0)h(δm0)2i.
(17)
Correspondingly, the cross susceptibility is the correlation of a single solute dipole with
the dipole moment δMs of the entire solvent in the sample [21]
χ0s = (β/3Ω0)hδm0 · δMsi.
(18)
Equation (17) neglects correlations between dipole moments of the solutes in the solution
represented by the corresponding Kirkwood factor. Since the latter describes short-range
correlations, of the length-scale of the molecular diameter [9], they can be safely omitted
in the type of theory developed here.
Both standard arguments of the dielectric theories [14] and microscopic derivation
[12] suggest a simple connection between the solute dipolar susceptibility χ0 and the
self susceptibility χ00
χ0 = χcχ00.
(19)
This relation implies that the account of the solute-solvent cross-correlations entering
susceptibility χ0s amounts to introducing the cavity field acting on the average solute
Protein dielectrics
7
dipoles, which also defines the torque acting on a selected dipole in the Onsager theory
of dipolar liquids (directing field) [13].
Adding the dipolar polarization of the solutes to equation (11) for the dielectric
constant of the liquid with spherical voids, one arrives at the dielectric constant of the
solution
ǫs
ǫmix
= 1 − 3η0 + 3η0ǫsχc (1 − y0) ,
(20)
where y0 = (4π/3)χ00. This equation clearly reduces to (11) in the limit of non-polar
solutes when y0 → 0.
2.3. Frequency-Dependent Response
The static arguments presented in the previous sections can be extended to the frequency
domain of main interest to broad-band dielectric spectroscopy. The dielectric constants
of both the solvent and the mixture become frequency-dependent functions, ǫs(ω) and
ǫmix(ω). The dipolar susceptibility of an isolated solute transforms into a linear response
function, instead of a static correlator of equation (17). The relevant formalism is
well developed and the result is the following response function of the solute dipolar
fluctuations [24, 25]
χ00(ω) = χ00h1 + iω S00(ω)i .
(21)
Here, S00(ω) is the Laplace-Fourier transform of the normalized time correlation function
of the solute dipole m0(t)
S00(t) = (cid:2)h(δm0)2i(cid:3)−1
hδm0(t) · δm0(0)i.
(22)
This function was fitted to multi-exponential decay when applied to the analysis of the
MD simulation data presented below
S00(t) = Xi
Aie−t/τi, Xi
Ai = 1,
(23)
where τi are the relaxation times and Ai are the relative weights of the relaxation
components. From this equation, one gets the frequency-dependent function y0(ω)
y0(ω) = y0Xi
Ai
1 − iωτi
.
The frequency-dependent dielectric constant of the solution becomes
ǫs(ω)
ǫmix(ω)
= 1 − 3η0 + 3η0ǫs(ω)χc(ω) (1 − y0(ω)) .
(24)
(25)
Our arguments so far have not included any approximations except neglecting
mutual polarization of solutes at their high concentration and the short-range
correlations of solute dipoles entering the Kirkwood factor of the solutes. However,
equations (20) and (25) contain an unknown cavity-field susceptibility χc(ω). The
Maxwell's result for this function refers to a free surface separating a dielectric from
It is a priory not obvious that this function can describe the complex and
a void.
Protein dielectrics
8
]
c
χ
[
e
R
0.25
0.2
0.15
0.1
0.05
110 mg/mL
28 mg/mL
57 mg/mL
111 mg/mL
water/buffer
0
0.01
0.1
1
ν, GHz
10
100
Figure 1: Real part of the cavity-field susceptibility χc(ω) extracted from experimental
dielectric measurements according to (25). The results combine the broad-band
dielectric measurements from [19] with high-frequency data from [20]. The dotted line
indicates χM
from equation (12) for pure water (at lower frequencies) and the buffer (at
c
higher frequencies). The gap between the two sets of curves represents the frequency
window between the measurements.
heterogeneous protein-water interface involving both weak protein-water interactions at
hydrophobic patches and strong binding to charged surface residues. However, one can
use the experimental input for the dielectric constants of the mixture and pure water in
equation (25) to extract the cavity-field susceptibility χc(ω).
c(ω) extracted from equation (25) using frequency-
Figure 1 shows the real part χ′
lysozyme solutions from broad-band dielectric
dependent dielectric constants of
spectroscopy below 50 GHz [19] and from separate measurements in the frequency range
70 -- 700 GHz [20]. The dotted line shows Re[χM
c ] from equation (12); the break in the
curve signals the transition from water to buffer used at higher frequencies in [20]. The
cavity-field susceptibility follows very closely the Maxwell prediction in the range of
frequencies 10 -- 200 GHz, but then deviates downward outside this range. The behavior
at low frequencies is particularly noteworthy.
It turns out that the dipole moment induced at the protein by an external field
is over-screened [26] by the hydration layers, and perhaps the ionic atmosphere, to
nearly zero. In fact, χc is below zero at ν < 1 GHz, implying a die-electric response,
i.e. repulsion of the protein dipole from a region of a stronger electric field. This
phenomenon, known as negative dielectrophoresis,
is well-documented for hydrated
nanoparticles [27], but has not been broadly observed for proteins. Our recent extensive
simulations of ubiquitin [12], which is neutral at pH= 7.0, have indicated exactly this
scenario: a negative χ0s, larger in magnitude than the positive χ00, thus resulting in
a slightly negative χ0 in equation (16). However, this result has not been detected by
simulations of charged proteins, including lysozyme, probably due to the neglect of the
ionic atmosphere in the analysis.
Figure 1 suggests that dielectric models of the cavity-field susceptibility do not
provide an adequate description in the entire range of frequencies of interest to broad-
Protein dielectrics
9
band spectroscopy. However, the modeling can proceed along separate routes since the
expulsion of polar water from the solute core is significant only at high frequencies,
while the polar response of the protein dipole, described by y0(ω), dominates at low
frequencies. One therefore can keep the Maxwell result for χc(ω) for the former
component, as realized in equation (9). Since there is currently no model allowing to
describe the overscreening observed at low frequencies, we have resorted to an empirical
approximation. Replacing χcǫsy0 in eqs (20) and (25) with χM
c y0 accomplishes most of
what is seen to occur in figure 1 and allows us to arrive at a compact relation for the
dielectric constant of the solution
ǫs(ω)
ǫmix(ω)
= 1 +
3η0
2ǫs(ω) + 1
[ǫs(ω) − 1 − 3y0(ω)] .
(26)
2.4. Dielectric instability
Equation (20) predicts a point of dielectric instability at which the assumption of a
uniform solution of weakly interacting protein dipoles breaks down. The instability
is toward clustering of dipoles and is associated with the divergence of the dielectric
constant ǫmix. It is reached at the critical volume fraction
3ηc = [1 + ǫsχc(y0 − 1)]−1 .
(27)
If the Maxwell form of the cavity-field susceptibility is used in this equation, the critical
point ηc = 0.01 (y0 ≃ 16) corresponds to the concentration of 8 mg/mL for lysozyme
in solution. Lysozyme solutions are stable in this range of concentrations and this
estimate is clearly too low. On the contrary, the overscreening scenario shown in figure
1 makes ηc negative, thus removing the instability altogether. While other forms of
aggregation are still possible [28, 29], it might be quite possible that overscreening of
the protein dipole eliminates instability toward dipolar clustering (such as formation
of dipolar chains) and lowers the sensitivity of proteins in solutions to inhomogeneous
electric fields always present in vivo.
3. Application to Experiment: Lysozyme Solution
Dielectric measurements of solutions typically provide the real and imaginary parts of the
dielectric constant as functions of frequency and solution composition [30, 31, 19, 32, 33].
The existence of these two coordinates, frequency and solute concentration, allows one to
learn about the specific pattern of interfacial polarization realized for a given solute and
the dynamics of processes contributing to the relaxation of the sample dipole moment.
We start with the analysis of the concentration dependence at a given frequency, followed
with the analysis of the frequency dependence at a fixed concentration.
3.1. Decrement of the water Debye peak
Independently of the details of the dynamics of a protein itself and its coupling to
the interfacial waters, the time-scales of these motions are significantly lower than the
Protein dielectrics
10
0
-0.05
s
ε
/
ε
∆
-0.1
-0.15
72 GHz
18 GHz
0
50
100
C, mg/mL
150
200
Figure 2: Decrement of the water dielectric constant in the solution of lysozyme in
water, ∆ǫ(ω)/ǫs(ω) = ǫmix(ω)/ǫs(ω) − 1, as a function of the protein concentration C.
The points are the experimental data at the frequency of the water Debye peak νD ≃ 18
GHz (circles) [19], and at ν = 72 GHz (diamonds) [20]. The solid and dash-dotted
lines refer to the calculations using equations (9) and (10) in the order of increasing
frequency. The dashed line is the calculation incorporating the dynamics of the protein
dipole according to equation (26) with y0(ωD) calculated from MD simulations [12]. The
lysozyme molecular volume of Ω0 = 29.8 nm3 is used to convert from the volume fraction
to the solution concentration. The dotted line connects the experimental points.
characteristic time of dielectric relaxation of water. The global motions of the solute
are dynamically frozen at the frequency of the water Debye peak (νD ∼ 18 GHz).
This implies that y0(ωD) can be dropped from equation (26). One then arrives at the
dielectric constant of the mixture of polar water and effectively non-polar solutes (eqs
(8) and (9)). Any sufficiently high frequency can in principle be taken for this analysis.
The decrement of the Debye peak of water in the solution vs the solute concentration
is often reported [8] and can be used, in the framework of the present theory, as a
convenient source of data to extract the information about the parameters α and χc.
Our formalism is applied to recent measurements of dielectric spectra of lysozyme
solutions [19, 20]. Figure 2 shows the dependence of the decrement in the amplitude
of the water Debye peak ωD in the solution ∆ǫ(ωD) = ǫmix(ωD) − ǫs(ωD) vs the protein
concentration. Circles show the experimental data from [19], while the solid and dashed
lines refer to equations (9) and (26), respectively. For the latter, y0(ωD) calculated from
MD simulations [12], and discussed below for the analysis at lower frequencies, was used.
Clearly, the protein permanent dipole can be safely neglected. The transformation from
the solution concentration reported experimentally to the volume fraction required by
(9) and (25) was performed by using the volume of lysozyme Ω0 = 29.8 nm3. The latter
was calculated from the crystallographic structure of the protein (3FE0, PBD database)
by using the algorithm developed by Till and Ullmann [34].
In accord with the results shown in figure 1, the cavity-field susceptibility is well
described by the Maxwell form (equation (12)) at the frequency of the water Debye
peak, and the agreement between theory and experiment is excellent. It becomes less
Protein dielectrics
11
(a)
]
)
ν
(
ε
[
e
R
80
60
40
20
110 mg/mL (exp)
28 mg/mL (exp)
water (exp)
28 mg/mL (calc)
110 mg/mL (calc)
0.001
0.01
0.1
1
10
100
(b)
]
)
ν
(
ε
[
m
I
10
1
0.1
0.001
0.01
0.1
ν, GHz
1
10
100
Figure 3: Real (a) and imaginary (b) parts of the dielectric constant of the lysozyme
solution measured experimentally [19] (points) and calculated theoretically (lines). The
dotted lines show the real and imaginary parts of the dielectric spectrum of water from
[36].
satisfactory at a higher frequency of 72 GHz [20], also shown in figure 2. The refractive
index of the protein starts to affect the result at this high frequency and np = 1.7 from
[35] was adopted in the calculations using equation (10). The theoretical slope with
increasing protein concentration is higher than experimentally reported and is likely
related to deviations from the Maxwell form of the cavity-field susceptibility seen in
figure 1.
3.2. Dielectric spectra of solutions
The results for the dispersion and loss spectra of lysozyme solutions are shown in
figure 3. Experimental data from [36] were used for ǫs(ω) and Molecular Dynamics
(MD) simulations of a single lysozyme protein hydrated in a simulation box of TIP3P
waters [12] were used to produce y0(ω) in (24). The relaxation parameters in (24) are:
Ai = {0.13, 0.06, 0.81}, τi = {0.037, 0.295, 14.6} ns, y0 = 16.3. The dominant relaxation
component of the solute dipole, with the relaxation time of 14.6 ns, can be assigned
to protein tumbling. The relaxation time of 9.1 ns was reported for this relaxation
component from the analysis of proton NMR at low resonance frequencies [37].
The usefulness of MD simulations is somewhat limited for the sake of comparison
with experiment since the charge distribution in the protein studied by simulations might
not entirely fit the experimental conditions. The standard force-field prescriptions for
protonating/deprotonating the surface residues of lysozyme at pH = 7.0 produce the
overall protein charge of +7, while the charge of +10 is reported at pH= 5.5 in the
Protein dielectrics
12
experimental study [19]. Overall, permanent dipole moments of proteins arise from
slight deviations from highly symmetric distribution of charge minimizing the total
dipole moment [38, 39]. Shifts of pKa values of surface residues due to local electrostatic
environment [40], ion association, and pH can therefore alter the dipole moment.
Despite remaining uncertainties regarding the magnitude of the protein dipole when
experimental conditions are concerned, the dipole hm0i = 223 D from MD results in
a fair agreement between theoretical and experimental dispersion curves ǫ′
mix(ω) at the
lower concentration of the protein, 28 mg/mL (figure 3a). The theory misses some of the
static dielectric constant at the higher concentration, c = 110 mg/mL, but the difference
actually comes from the missing increment at intermediate frequencies associated with
δ-dispersion. This part of the spectrum is also missing in the loss spectrum (figure 3b).
This outcome is expected since specific protein-water binding contributing to this signal
[21, 41, 30] has not been incorporated into the model.
4. Summary
Broad-band dielectric spectroscopy is a widely used tool to interrogate the dynamics
of complex systems, including protein solutions. The interest in the field in the recent
years has been to extract the polarization properties and dynamics of protein hydration
layers from frequency-dependent spectra. The standard approach to the problem is to
fit the dispersion and loss spectra to a sum of Debye or stretch-exponential functions,
assuming that each component represents a separate relaxation process in a complex
environment. The obvious limitation of this approach is the non-additivity of interfacial
polarization, well recognized by classical theories of dielectric mixtures [2, 3, 4]. While
these classical theories were developed for mixtures of dielectric materials, when each
component can be represented by a macroscopic dielectric body, their application to
hydrated proteins is clearly limited. At the same time, standard theories of dipolar
liquids [10, 9] are not of much use either since they do not recognize the existence of
an extended polarizable interface, which is in fact the central concept of the dielectric
theories of mixtures. The present theoretical development aims to fill the void existing
in each approach by recognizing both the molecular nature of the protein dipole and a
quasi-macroscopic subensemble of interfacial waters producing interfacial polarization.
The theory thus aims to study if the standard rules established for cavities carved
in dielectrics, and also applied to calculate the local field acting on molecular dipoles
[13], can be applied to hydrated proteins. Equations (25) is central to this analysis
since it allows us to extract the cavity-field susceptibility χc(ω) directly from the
frequency-dependent dielectric constants of the protein solution and pure water. The
remarkable result of this analysis is that at ω < 1 GHz the susceptibility χc(ω) is below
≃ 0.02 predicted by the Maxwell equation (12) and is in the negative territory, down to
≃ −10−3. Therefore, the standard prescription derived for dielectric cavities (equations
(1) and (12)) cannot be used in successful theories of dielectric response of protein
solutions.
Protein dielectrics
13
Granted, the cavity susceptibility extracted from experimental measurements might
reflect the combined response of the dielectric interface and the ionic atmosphere.
However, as a cumulative signature of the protein-water interface,
it dramatically
downscales the permanent dipole sensed by the dielectric experiment compared to its
value calculated from atomic charges. Its low value can also help to explain the puzzling
ability of proteins to stay in solution in vivo, despite significant electric field gradients
that should pull a paraelectric particle to stick to, for instance, the bilipid membrane.
The die-electric response suggested by the present analysis of experimental data, and
our previous simulations [12], might be an answer to this puzzle since a die-electric
solute repels from a charged interface creating the field gradient. It also eliminates the
dielectric instability toward clustering of the solute dipoles predicted by (20) and (27)
when the Maxwell form of the cavity-field susceptibility is used there.
Acknowledgments
This research was supported by the National Science Foundation (DVM, CHE-0910905).
CPU time was provided by the National Science Foundation through TeraGrid resources
(TG-MCB080116N). The author is grateful to Drs. Cametti and Nguyen for sharing their
experimental results.
References
[1] C. J. F. Bottcher and P. Bordewijk. Theory of electric polarization. Deielctrics in time-dependent
fields, volume 2. Elsevier, Amsterdam, 1978.
[2] S. Takashima. Electrical properties of biopolymers and membranes. Adam Hilger, Bristol, 1989.
[3] B. K. P. Scaife. Principles of dielectrics. Clarendon Press, Oxford, 1998.
[4] T. C. Choi. Effective Medium Theory. Clarendon Press, Oxford, 1999.
[5] P. E. Smith, R. M. Brunne, A. E. Mark, and W. F. van Gunsteren. Dielectric properties of
trypsin inhibitor and lysozyme calculated from molecular dynamics simulations. J. Phys. Chem.,
97:2009 -- 2014, 1993.
[6] J. W. Pitera, M. Falta, and W. F. van Gunsteren. Dielectric properties of proteins from
simulations: The effects of solvent, ligands, ph, and temperature. Biophys. J., 80:2546, 2001.
[7] G. King, F. S. Lee, and A. Warshel. Microscopic simulations of macroscopic dielectric constant of
solvated proteins. J. Chem. Phys., 95:4366, 1991.
[8] G. P. South and E. H. Grant. Dielectric dispersion and dipole moment of myoglobin in water.
Proc. R. Soc. Lond. A, 328:371 -- 387, 1972.
[9] G. Stell, G. N. Patey, and J. S. Høye. Dielectric constants of fluid models: Statistical mechanical
theory and its quantitayive implementation. Adv. Chem. Phys., 48:183, 1981.
[10] P. Madden and D. Kivelson. A consistent molecular treatment of dielectric phenomena. Adv.
Chem. Phys., 56:467, 1984.
[11] A. D. Friesen and D. V. Matyushov. Local polarity excess at the interface of water with a nonpolar
solute. Chem. Phys. Lett., 511:256 -- 261, 2011.
[12] D. V. Matyushov. Dipolar response of hydrated proteins. J. Chem. Phys., 136:085102, 2012.
[13] L. Onsager. Electric moments of molecules in liquids. J. Am. Chem. Soc., 58:1486, 1936.
[14] C. J. F. Bottcher. Theory of Electric Polarization, volume 1. Elsevier, Amsterdam, 1973.
[15] J. D. Jackson. Classical electrodynamics. Wiley, 1999.
Protein dielectrics
14
[16] J. S. Høye and G. Stell. Statistical mechanics of polar systems. II. J. Chem. Phys., 64(5):1952 --
1966, 1976.
[17] D. R. Martin and D. V. Matyushov. Cavity field in liquid dielectrics. Europhys. Lett., 82:16003,
2008.
[18] D. R. Martin, A. D. Friesen, and D. V. Matyushov. Electric field inside a "Rossky cavity" in
uniformly polarized water. J. Chem. Phys., 135:084514, 2011.
[19] C. Cametti, S. Marchetti, C. M. C. Gambi, and G. Onori. Dielectric relaxation spectroscopy
of lysozyme aqueous solutions: Analysis of the delta-dispersion and the contribution of the
hydration water. J. Phys. Chem. B, 115(21):7144 -- 7153, 2011.
[20] N. Q. Vinh, S. James Allen, and Kevin W. Plaxco. Dielectric spectroscopy of proteins as a
quantitative experimental test of computational models of their low-frequency harmonic motions.
J. Am. Chem. Soc., 133(23):8942 -- 8947, 05 2011.
[21] S. Boresch, P. Hochtl, and O. Steinhauser. Studying the dielectric properties of a protein solution
by computer simulation. J. Phys. Chem. B, 104:8743, 2000.
[22] D. V. Matyushov. Terrahertz response of dielectric mixtures. Phys. Rev. E, 81:021914, 2010.
[23] H. Frohlich. Theory of dielectrics. Oxford University Press, Oxford, 1958.
[24] J. P. Hansen and I. R. McDonald. Theory of Simple Liquids. Academic Press, Amsterdam, 2003.
[25] Nilashis Nandi and Biman Bagchi. Anomalous dielectric relaxation of aqueous protein solutions.
J. Phys. Chem. A, 102(43):8217 -- 8221, 1998.
[26] V. Ballenegger and J.-P. Hansen. Dielectric permittivity profiles of confined polar liquids. J.
Chem. Phys., 122:114711, 2005.
[27] T. B. Jones. Electromechanics of Particles. Cambridge University Press, Cambridge, 1995.
[28] Yun Liu, Lionel Porcar, Jinhong Chen, Wei-Ren Chen, Peter Falus, Antonio Faraone, Emiliano
Fratini, Kunlun Hong, and Piero Baglioni. Lysozyme protein solution with an intermediate
range order structure. J. Phys. Chem. B, 115(22):7238 -- 7247, 2010.
[29] Fr´ed´eric Cardinaux, Emanuela Zaccarelli, Anna Stradner, Saskia Bucciarelli, Bela Farago,
Stefan U. Egelhaaf, Francesco Sciortino, and Peter Schurtenberger. Cluster-driven dynamical
arrest in concentrated lysozyme solutions. J. Phys. Chem. B, 115(22):7227 -- 7237, 2011.
[30] A. Oleinikova, P. Sasisanker, and H. Weingartner. What can really be learned from dielectric
spectroscopy of protein solutions? a case study of Ribonuclease A. J. Phys. Chem. B, 108:8467,
2004.
[31] Hermann Weingartner, Andrea Knocks, Stefan Boresch, Peter Hochtl, and Othmar Steinhauser.
Dielectric spectroscopy in aqueous solutions of oligosaccharides: Experiment meets simulation.
J. Chem. Phys., 115:1463 -- 1472, 2001.
[32] K.-J. Tielrooij, J. Hunger, R. Buchner, M. Bonn, and H. J. Bakker. Influence of concentration and
temperature on the dynamics of water in the hydrophobic hydration shell of tetramethylurea.
J. Am. Chem. Soc., 1332:15671 -- 15678, 2010.
[33] Hafiz M. A. Rahman, Glenn Hefter, and Richard Buchner. Hydration of formate and acetate ions
by dielectric relaxation spectroscopy. J. Phys. Chem. B, page 10.1021/jp207504d, 2011.
[34] Mirco S. Till and G. Matthias Ullmann. Mcvol - a program for calculating protein volumes and
identifying cavities by a monte carlo algorithm. J. Mol. Mod., 16:419, 2010.
[35] J. Knab, J.-Y. Chen, and A. Markelz. Hydration dependence of conformational dielectric
relaxation of lysozyme. Biophys. J., 90:2576 -- 2581, 2006.
[36] H. Yada, M. Nagai, and K. Tanaka. The intermolecular stretching vibration mode in water isotopes
investigated with broadband terahertz time-domain spectroscopy. Chem. Phys. Lett., 473:279,
2009.
[37] A. Krushelnitsky.
Intermolecular electrostatic interactions and Brownian tumbling in protein
solutions. Phys. Chem. Chem. Phys., 8:2117 -- 2128, 2006.
[38] D. J. Barlow and J. M. Thornton. Charge distribution in proteins. Biopolymers, 25:1717 -- 1733,
1986.
[39] S. Takashima. Electric dipole moment of globular proteins: measurement and calculation with
Protein dielectrics
15
nmr and x-ray databases. J. Noncrystal. Solids, 305:303 -- 310, 2002.
[40] B. L. Mellor, S. Khadka, D. D. Busath, and B. A. Mazzeo. Influence of pka shifts on the calculated
dipole moments of proteins. Protein J., 30:490 -- 498, 2011.
[41] Hermann Weingartner, Andrea Knocks, Stefan Boresch, Peter Hochtl, and Othmar Steinhauser.
Dielectric spectroscopy in aqueous solutions of oligosaccharides: Experiment meets simulation.
J. Chem. Phys., 115(3):1463 -- 1472, 2001.
|
1208.2837 | 1 | 1208 | 2012-08-14T11:52:56 | Glucose oxidase immobilization onto carbon nanotube networking | [
"physics.bio-ph"
] | When elaborating the biosensor based on single-walled carbon nanotubes (SWNTs), it is necessary to solve such an important problem as the immobilization of a target biomolecule on the nanotube surface. In this work, the enzyme (glucose oxidase (GOX)) was immobilized on the surface of a nanotube network, which was created by the deposition of nanotubes from their solution in 1,2-dichlorobenzene by the spray method. 1-Pyrenebutanoic acid succinimide ester (PSE) was used to form the molecular interface, the bifunctional molecule of which provides the covalent binding with the enzyme shell, and its other part (pyrene) is adsorbed onto the nanotube surface. First, the usage of such a molecular interface leaves out the direct adsorption of the enzyme (in this case, its activity decreases) onto the nanotube surface, and, second, it ensures the enzyme localization near the nanotube. The comparison of the resonance Raman (RR) spectrum of pristine nanotubes with their spectrum in the PSE environment evidences the creation of a nanohybrid formed by an SWNT with a PSE molecule which provides the further enzyme immobilization. As the RR spectrum of an SWNT:PSE:GOX film does not essentially differ from that of SWNT:PSE ones, this indicates that the molecular interface (PSE) isolates the enzyme from nanotubes strongly enough. The efficient immobilization of GOX along the carbon nanotubes due to PSE is confirmed with atom-force microscopy images. The method of molecular dynamics allowed us to establish the structures of SWNT:PSE:GOX created in the aqueous environment and to determine the interaction energy between hybrid components. In addition, the conductivity of the SWNT network with adsorbed PSE and GOX molecules is studied. | physics.bio-ph | physics |
V.A. KARACHEVTSEV, A.YU. GLAMAZDA, E.S. ZARUDNEV et al.
GLUCOSE OXIDASE IMMOBILIZATION ONTO CARBON
NANOTUBE NETWORKING
V.A. KARACHEVTSEV,1 A.YU. GLAMAZDA,1 E.S. ZARUDNEV,1
M.V. KARACHEVTSEV,1 V.S. LEONTIEV,1 A.S. LINNIK,1 O.S. LYTVYN,2
A.M. PLOKHOTNICHENKO,1 S.G. STEPANIAN1
1B. Verkin Institute for Low Temperature Physics and Engineering,
Nat. Acad. of Sci. of Ukraine
(47, Lenin Ave., Kharkov 61103, Ukraine)
PACS 78.67.Ch
c(cid:13)2012
2V. Lashkaryov Institute of Semiconductor Physics, Nat. Acad. of Sci. of Ukraine
(41 Nayki Ave., 03028 Kyiv, Ukraine)
When elaborating the biosensor based on single-walled carbon nan-
otubes (SWNTs), it is necessary to solve such an important prob-
lem as the immobilization of a target biomolecule on the nan-
otube surface. In this work, the enzyme (glucose oxidase (GOX))
was immobilized on the surface of a nanotube network, which was
created by the deposition of nanotubes from their solution in 1,2-
dichlorobenzene by the spray method. 1-Pyrenebutanoic acid suc-
cinimide ester (PSE) was used to form the molecular interface,
the bifunctional molecule of which provides the covalent binding
with the enzyme shell, and its other part (pyrene) is adsorbed
onto the nanotube surface. First, the usage of such a molecular
interface leaves out the direct adsorption of the enzyme (in this
case, its activity decreases) onto the nanotube surface, and, sec-
ond, it ensures the enzyme localization near the nanotube. The
comparison of the resonance Raman (RR) spectrum of pristine
nanotubes with their spectrum in the PSE environment evidences
the creation of a nanohybrid formed by an SWNT with a PSE
molecule which provides the further enzyme immobilization. As
the RR spectrum of an SWNT:PSE:GOX film does not essentially
differ from that of SWNT:PSE ones, this indicates that the molec-
ular interface (PSE) isolates the enzyme from nanotubes strongly
enough. The efficient immobilization of GOX along the carbon
nanotubes due to PSE is confirmed with atom-force microscopy
images. The method of molecular dynamics allowed us to establish
the structures of SWNT:PSE:GOX created in the aqueous envi-
ronment and to determine the interaction energy between hybrid
components. In addition, the conductivity of the SWNT network
with adsorbed PSE and GOX molecules is studied. The adsorp-
tion of PSE molecules onto the SWNT network causes a decrease
of the conductivity, which can be explained by the appearance
of scattering centers for charge carriers on the nanotube surface,
which are created by PSE molecules.
1. Introduction
Due to their unusual physical, optical, thermal, and
electronic properties, single-walled carbon nanotubes
(SWNTs) have a huge potential of different promising
applications including the biosensing. The main chal-
lenge in the development of the devices is the biofunc-
tionalization of nanomaterial surfaces and the creation
of appropriate interfaces between the nanotubes and the
biosystems. Carbon-nanotube-based biological sensors
could find applications, for example, in measuring the
concentrations of glucose in blood. This is particularly
important because the number of diabetics in the world
increases continuously and dramatically.
In the last few years, the applications of conducting
properties of carbon nanotubes in the biological sensor-
ing have demonstrated the high efficiency of this ap-
proach, despite the elaborations being in their infancy
[1 -- 3]. The use of the conductivity of carbon nanotubes
as the detection and measuring method in midget biosen-
sors showed some preferences over the optical methods
currently being employed in the clinical work. This is
related to the fact that the most of biological processes
involve electrostatic interactions and/or a charge trans-
fer, which can be directly detected with the electronic
equipment. Carbon nanotubes can be easily integrated
into the electronic device. Moreover, the average size
of an nanotube is usually compatible with the molecu-
lar size of a compound being analyzed, which increases
the sensitivity of the measurement. The conductivities
of individual nanotubes or carbon nanotube networks
can be utilized in the investigation and the develop-
ment of biosensors [4, 5]. Networks formed by hundreds
or thousands of carbon nanotubes distributed randomly
between metallic contacts have a larger active area for
the detection and can operate at higher currents. The
technology for the fabrication of carbon-nanotube net-
works is well developed and does not require the ex-
pensive equipment to operate. The networks can be
produced from solutions by deposition onto various sub-
strates.
700
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
GLUCOSE OXIDASE IMMOBILIZATION ONTO CARBON NANOTUBE NETWORKING
When working out biological sensors,
in which an
SWNT network is used, it is necessary to solve such an
important problem as the immobilization of a recogni-
tion biomolecule on the nanotube surface. In researches,
an enzyme immobilized on a nanotube is often used as
the recognition element.
In some works, the enzyme
was immobilized directly onto the nanotube surface [6].
However, the recent work [7] showed that the activity
of two enzymes (R-chymotrypsin and soybean peroxi-
dase) decreased significantly after their adsorption onto
the surface of single-walled carbon nanotubes. Thus, the
problem of enzyme immobilization on a nanotube that
needs to be solved is to retain the enzyme native activity
despite its immobilization on the nanotube surface. The
problem is closely related to the nanotube functional-
ization. Some success on this front has been achieved
owing to the use of a molecular anchor [8, 9] or the
polymer-wrapping of nanotubes [6]. The efficient non-
covalent carbon nanotube functionalization by organic
molecules for the biocompatibility testing was suggested
by Chen and co-workers [8]. Their approach utilizes a bi-
functional molecule containing succinimidyl ester and a
pyrene moiety to bind proteins to the nanotube surface.
Pyrene attaches to the nanotube surface by means of the
π − π stacking and the hydrophobic interaction and does
not disrupt the nanotube backbone. Another fragment
of the anchor molecule is succinimidyl ester that binds
enzymes to the nanotube surface. This molecular in-
terlayer provides a sufficiently strong attachment of the
enzyme to the nanotube surface and leaves the enzyme
activity unaffected. The enzyme localization near the
nanotube needs to provide the reliable detection of the
charge, which appears as a result of the biochemical re-
action of enzymes with the probe.
In the present work, the enzyme (glucose oxidase
(GOX)) was immobilized onto the surface of a nanotube
network, which was created by the deposition of nan-
otubes from their solution in dichlorbenzene by the spray
method. 1-pyrenebutanoic acid N-hydroxysuccinimide
ester (PSE) was used to form the molecular interface.
The comparison of the resonance Raman (RR) light scat-
tering spectrum of pristine nanotubes with their spec-
trum in the PSE environment evidences the creation of
a nanohybrid formed by SWNT with a PSE molecule
which ensures the further enzyme immobilization. As
the RR spectrum of an SWNT:PSE:GOX film does not
essentially differ from that of SWNT:PSE ones, this in-
dicates that the molecular interface (PSE) isolates the
enzyme from nanotubes efficiently. The immobilization
of GOX near a carbon nanotube due to PSE is con-
firmed by atom-force microscopy (AFM). The method
of molecular dynamics allows us to establish the struc-
tures of SWNT:PSE:GOX and to determine the energies
of the intermolecular interaction between components of
the triple complex in the aqueous environment. In ad-
dition, we study the conductivity of the SWNT network
with adsorbed PSE and GOX molecules. The adsorp-
tion of PSE molecules on the SWNT network results in
a decrease of the conductivity, which is most likely in-
duced by the appearance of scattering centers for charge
carriers in nanotubes.
2. Experimental
SWNTs have been produced by the CoMoCAT method
(SouthWest NanoTechnologies Inc., USA). This method
yields nanotubes with narrow diameters (0.75 -- 0.95 nm)
and chirality (6,5) [10]. PSE was purchased from Sigma-
Aldrich, Europe. All compounds have been used without
additional purification.
Nanotube suspensions were prepared in dichloroben-
zene (Sigma-Aldrich, Europe), which is the most effi-
cient organic solvent for nanotubes. Suspensions were
obtained by ultrasonication (44 kHz, UZDN-2, Sumy,
Ukraine) and by the following ultracentrifugation (up
to 18 000 g). As a result, all insoluble thick bundles of
nanotubes, as well as the metallic catalyst, were precipi-
tated. To get one sample, we used 0.2 mg of nanotubes,
being dissolved in 4 ml of dichlorobenzene.
The carbon nanotube network on the substrate
(quartz) was created by the spray method. To form
this network, a special electronic device was constructed
to control the nanotube network density, by using the
conductivity of the deposited network and a quartz mi-
crobalance. This density was regulated by the variation
of the deposition time, as well as with the nanotube con-
centration in a solution. After the drying of such a nan-
otube network, two gold contact areas with 10 µm gap
between them were thermodeposited. Glucose oxidase
(GOX) was immobilized onto the carbon nanotube net-
work through PSE, which formed a molecular interface
between nanotubes and the enzyme [8]. PSE was ad-
sorbed from its solution in methanol (10−4 M), and then
GOX from its solution in water (10−4 M) was deposited.
The GOX immobilization onto carbon nanotubes was
controlled with an AFM Nanoscope D3000 (Digital In-
struments,USA), as well as with RR spectroscopy. For
AFM measurements, nanotubes were sprayed on mica.
For Raman measurements, three dried films were pre-
pared: pristine nanotubes, nanotubes with PSE, and
SWNT:PSE:GOX. To prepare the film of nanotubes with
adsorbed PSE molecules, SWNTs were mixed with PSE
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
701
V.A. KARACHEVTSEV, A.YU. GLAMAZDA, E.S. ZARUDNEV et al.
Fig. 1. AFM image of the carbon nanotube network on a mica
substrate obtained by the spray method from an SWNT solution
in dichlorobenzene
in methanol (with 1:1 weight ratio, 0.3 mg/ml). The
mixture was treated with the sonication (1 W, 44 kHz)
for 30 min. Then the suspension was deposited on the
quartz substrate and dried under a stream of warm
air. SWNT:PSE:GOX hybrids were prepared by the
adsorption of GOX from an aqueous solution on an
SWNT:PSE film. Raman experiments have been per-
formed in the 90◦ scattering configuration relative to
the laser beam, by using 632.8 nm (1.96 eV) light from
a He-Ne laser. The spectra were analyzed using a Ra-
man double monochromator with the reverse dispersion
of 3A/mm and were detected with a thermocooled CCD
camera. The position of peaks of the bands in the RR
spectrum of a nanotube film was determined with the ac-
curacy not worse than 0.3 cm−1. This level of accuracy
has been achieved due to the observation of the plasma
lines of a laser in the spectra in a vicinity of the bands
corresponding to the tangential mode (G) and the radial
breathing mode (RBM) of nanotubes, which were used
in the internal calibration of a spectrometer.
Volt-ampere characteristics of the nanotube network
were determined by the home-built set-up based on a
microcontroller, by means of which the required range
of voltages was formed. A microamperemeter, which
measured a current through the network, was connected
with a computer. A minimal step of the output voltage
was 2 mV, and the current sensitivity reached 0.1 nA.
SWNT hybrids with PSE and with GOX were mod-
eled by the molecular dynamics method employing
NAMD programs [11]. In these calculations, the force
field Charmm27 was used [12]. For the enzyme, the stan-
dard force parameters were applied, and the force param-
eters of aromatic carbon were given to carbon nanotube
atoms. Glucose oxidase was obtained from the Protein
Data Bank. Its structure parameters were determined
earlier [13]. A PSE molecule has no standard parameters
in this force field. Therefore, an additional calculation
within the DFT method was made (PSE parametriza-
a
b
Fig. 2.
substrate
AFM image of SWNT:PSE:GOX hybrids on a mica
tion is described in details in [14]). In this work, carbon
nanotubes 7.95 A in diameter, 95.2 A in length, and with
chirality (10,0) were used. Each system was placed in a
cubic water box. The distance between the hybrid un-
der study and walls of the water box was not less than
12 A. Upon modeling, periodic conditions were used.
Upon modeling the pressure (1 bar), the temperature
(293 ◦K) and the atom number in the system were un-
changed. At the modeling beginning, the energy of the
system was minimized during 1000 cycles. The VMD
Program was applied for the visualization of the results
of calculations [15].
3. Results and Discussion
3.1. Analysis of the enzyme immobilization
onto SWNTs by AFM
We used the AFM method to study the morphology of
the network of SWNT deposited onto mica from a so-
lution of nanotubes in dichlorobenzene. Figure 1 shows
the network of SWNTs obtained by the spray method
on mica. Both single nanotubes and their bundles can
702
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
GLUCOSE OXIDASE IMMOBILIZATION ONTO CARBON NANOTUBE NETWORKING
Fig. 3.
Raman spectra of SWNTs (1), SWNT:PSE (2)
SWNT:PSE:GOX (3) hybrids in the range of the RBM and G
modes. Each experimental spectrum obtained by the laser excita-
tion at λexc = 632.8 nm
Fig. 4. Low-frequency fragment of the spectra of SWNTs (bold
line), SWNT:PSE (thin line), and SWNT:PSE:GOX (broken line)
in the RBM range
be seen in Fig. 1. Then we deposited the molecular in-
terface (PSE) on the network, and finally the enzyme
GOX was immobilized. The AFM image of these hy-
brids is presented in Fig. 2. As is seen from Fig. 2, the
GOX globular structures are mainly placed along nan-
otubes, and their heights are in the 4.38 -- 6.37-nm range.
As the height of one enzyme globule is about 4 nm [6],
the obtained heights are the sum of heights of a COX
globule and a nanotube. The surface of some nanotubes
is fully covered with the enzyme. The detailed analysis
of hybrid heights revealed that they may be 8 -- 8.5 nm
in some cases, and this indicates that the enzyme dimer
is created on a nanotube. Thus, it may be concluded
that, due to the PSE molecular interface, the enzyme is
placed near the nanotube surface.
3.2. Raman spectroscopy of carbon nanotubes
in the PSE and GOX environments
Information on the noncovalent interactions of organic
molecules with the nanotube surface can be obtained
from the analysis of the RR spectra of carbon nanotubes
before and after the deposition of these molecules [16].
The noncovalent interaction of a nanotube and an or-
ganic molecule results in a shift of the nanotube vibra-
tional modes and in the intensity redistribution between
bands. At first, we studied such spectra of pristine nan-
otubes in bundles. Then we measured the spectra of
nanotubes with deposited PSE molecules and, finally,
after the immobilization of GOX molecules. Figure 3
presents the RR spectra of a nanotube network in the
225 -- 1650 cm−1 range before (1) and after (2) the de-
position of PSE, as well as after the addition of GOX
(3). It should be noted that the spectra are similar, but
the detailed analysis of some bands reveals some slight
change in the spectrum after the PSE deposition onto
the nanotube surface.
To analyze these changes in more details, we studied
the most informative fragments of the RR spectra of
nanotubes: ranges of the RBM (225 -- 350 cm−1) and G
(1500 -- 1650 cm−1) modes.
3.2.1 Radial breathing mode
Figure 4 presents the low-frequency fragment of the RR
spectrum of nanotubes, in which the RBM is observed.
At the excitation by a He -- Ne laser in the 225 -- 350 cm−1
range, seven intense bands are observed for nanotubes
obtained by the CoMoCat method. The spectra were
approximated with a sum of seven Lorentzians. The fre-
quency position of the peaks, area, and width at the
half of the height of bands for the both samples were
determined. These results are presented in Table 1. The
nanotube chirality shown in Table 1 was determined ear-
lier for these nanotubes [17]. As follows from Table 1, all
bands were assigned to semiconducting nanotubes. The
spectra were normalized to the intensity of the high-
frequency component of the G-mode (G+), which al-
lowed us to compare the band intensities in the spectra
of three samples.
As follows from Table 1, the spectral shift in the RBM
(∼1.4 cm−1) is observed for SWNT:PSE in comparison
with SWNT bundles. This shift is caused by the in-
teraction of PSE molecules with nanotubes.
In addi-
tion, a decrease and an enhancement of the band inten-
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
703
V.A. KARACHEVTSEV, A.YU. GLAMAZDA, E.S. ZARUDNEV et al.
T a b l e 1. Position (ωRBM, cm−1), width at the half of the height (∆Γ, cm−1), integral intensity of Lorentzian
curves normalized to the band corresponding to the G+ band, chirality of nanotubes (n,m), and second electronic
transition of the semiconducting SWNTs (ES
22, eV) located close to the excitation energy of a He -- Ne laser (1.96 eV)
(n,m)
E S
22, eV
[17]
ω, cm−1
10.3
11.1
7.6
7.5
8.3
6.5
6.4
1.90
2.00
1.90
1.88
1.87
2.18
2.09
252.4
257.9
265.5
284.2
298.1
310.7
337.2
SWNT
∆Γ,
cm−1
7.4
8.6
7.4
7.1
9.6
7.2
4.9
S/S+
G
×100%
ω, cm−1
0.24
0.56
0.37
2.49
1.23
0.26
0.21
251.4
257.9
264.1
283.6
297.3
310.8
336.7
SWNT:PSE
SWNT:PSE:GOX
∆Γ,
cm−1
7.2
7.0
6.9
7.2
9.4
8.3
5.2
S/S+
G
×100%
0.25
0.40
0.22
1.36
0.59
0.21
0.29
ω,
cm−1
252.4
257.9
265.8
284.7
299.2
311.0
336.6
∆Γ,
cm−1
5.8
7.4
7.1
6.7
8.8
6.9
4.7
S/S+
G
×100%
0.12
0.59
0.19
1.22
0.36
0.19
0.37
a
b
Fig. 5.
Scheme of the resonant windows of semiconducting nanotubes with different chiralities (6,4) (a) and (7,5) (b) in the SDS
surrounding (solid curve), SWNTs in the film (thin line) and SWNT:PSE (broken line). The energy (1.96 eV) of a He-Ne laser is shown
by the vertical line. Lorentzian curves were used to describe the contours of resonant windows
sities are seen at 265.5, 284.2, 298.1 cm−1 and at 257.9,
310.7, 337.2 cm−1, respectively.
It is known that the
RBM bands are very sensitive to resonance conditions,
which are determined by the location of the nanotube
electronic level relative to the laser energy.
It is well
known that the position of the nanotube electronic level
depends on the interaction with the environment [18].
Thus, when the nanotube electronic level (in our case,
it is the second electronic level (E S
22)) is located higher
than the energy of a laser, it becomes lower upon the
environmental interaction, the resonance conditions are
improved, and the intensity of the corresponding band
increases. Otherwise, the resonance conditions become
worse, and the intensity of the corresponding band de-
creases.
The resonance windows for two semiconducting nan-
otubes with different positions of electronic levels rel-
atively to the laser energy (1.96 eV) are schematically
shown in Fig. 5. The laser energy is marked with the ver-
tical line. Other lines are obtained using Lorentz func-
tions. In addition, the bands corresponding to the aque-
ous suspension of nanotubes with a surfactant (sodium
dodecyl sulfate (SDS)) (thick line) and in the film (thin
line) and their hybrids with PSE (thin broken line) are
shown in Fig. 5. The energy of the electronic transi-
tion for nanotubes (E S
22) in the aqueous suspension of
individual nanotubes with the SDS is by 60 -- 160 meV
higher, than the energy of this level for nanotubes in
a film [18, 19]. This red shift is due to the strong in-
teraction of nanotubes in bundles, which are situated
in a film. For SWNT:PSE, the bundling effect is much
smaller, because the organic molecules become embed-
ded between nanotubes in bundles in a solution during
the ultrasonication and, thus, decrease the interaction of
nanotubes. As a result, due to the interaction between
nanotubes and organic molecules, the E S
22 level for such
704
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
GLUCOSE OXIDASE IMMOBILIZATION ONTO CARBON NANOTUBE NETWORKING
SWNTs is still lower than that for individual nanotubes
in an aqueous suspension.
In our case, the nanotubes with (11,1), (6,5), and (6,4)
chiralities have the electronic level (E S
22), which is higher
than the laser energy. So, under the interaction of nan-
otubes with PSE, the band intensity increases. This rule
works unambiguously for nanotubes with (6,4) chiral-
ity. Contrary to this, the level E S
22 of nanotubes with
(10,3), (7,6), (7,5), and (8,3) chiralities is lower than the
laser energy. Therefore, under the interaction of these
nanotubes with PSE, the band intensity must decrease,
which is observed experimentally.
We note that, after the adsorption of GOX molecules
onto SWNT:PSE, the spectral transformation is weaker.
For example, the intensities of bands assigned to (7,5),
(10,3) and (7,6) nanotubes do not change practically.
3.2.2 Tangential mode
The most intense band in the RR spectra of SWNTs
lies in the 1550 -- 1620 cm−1 range and corresponds to
the high-frequency component of the tangential G mode
[20]. This band for nanotubes and their hybrids with
PSE and PSE:GOX is shown in Fig. 6. As is seen
from Fig.
6, the intense band in the RR spectrum
of the nanotubes in bundles (1592.4 cm−1) is low-
shifted by 1.7 cm−1 in the SWNT:PSE film. This
shift takes place, because the interaction energy of nan-
otubes in bundles decreases due to a weaker nanotube-
PSE binding energy. The intense band in the spec-
trum of the SWNT:PSE:GOX complex has a peak at
1590.5 cm−1, which indicates the insignificant influence
of GOX on the RR spectrum of nanotubes:PSE (located
at 1590.7 cm−1).
Some differences in the band positions and intensi-
ties were obtained after the fitting with the sum of
approximation functions. Each experimental spectrum
has been fitted with a minimal number of approxima-
tion functions (Fig. 6) consisting of the sum of four
Lorentzians and one Breit -- Wigner -- Fano (BWF) func-
tion, I(ω) = I0{1+(ω−ω0)/qΓ}2/{1+[(ω−ω0)/Γ]2} [21],
where I0, ω0, Γ, and q are the intensity, BWF frequency,
broadening parameter, and asymmetry parameter, re-
spectively. The BWF function is used for describing the
low-frequency spectral band, which appears due to the
presence of metallic nanotubes. On the basis of a good
coincidence between the experimental spectrum and the
sum of approximating curves, the parameters of these
curves were determined (Table 2). The total intensity of
the whole spectral fragment was set as 100%. As is seen
from Table 2, a low-frequency shift from 0.3 to 3.4 cm−1
Fig. 6.
High-frequency fragment of SWNTs (bold line),
SWNT:PSE (thin line), and SWNT:PSE:GOX (broken line) hy-
brids in the G mode range
range was observed in the film with PSE for all bands of
nanotubes.
Thus, in accordance with changes in the RR spec-
tra of nanotubes in the PSE environment relative to
the spectrum of pristine nanotubes, we can conclude
that the PSE molecule forms a noncovalent hybrid with
a carbon nanotube, which favors the further immobi-
lization of enzyme molecules. At the same time, the
absence of significant changes in the RR spectrum of
the SWNT:PSE:GOX hybrid (as compared with the
SWNT:PSE one) evidences that these PSE molecules
isolate efficiently the GOX molecules from nanotubes,
preventing their direct interaction with the nanotube
surface.
3.3. Structure and interaction energy of the
SWNT:PSE:GOX complex in the water
environment: molecular dynamics modeling
While modeling the SWNT:PSE:GOX complex, we were
aimed to reveal whether one PSE molecule is able to
keep one GOX molecule (covalently bound to it) near
the nanotube surface. The other task was to determine
the interaction energies between the PSE:GOX complex
and SWNTs. It is known that PSE reacts with amino
groups of peptides [8[.
In this case, the CO-NH bond
is formed. One PSE molecule was attached to one of
the lysine side residues of GOX. As a result, a new com-
pound was obtained (denoted as PSE:GOX). It consists
of the pyrene fragment and a GOX molecule joined with
a linker. The PSE:GOX complex was connected to the
SWNT surface by the pyrene fragment at a distance of
3.5 A. The system was minimized during 1000 cycles
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
705
V.A. KARACHEVTSEV, A.YU. GLAMAZDA, E.S. ZARUDNEV et al.
T a b l e 2. Position (ωRBM, cm−1), width at the half of the height (∆Γ, cm−1), and integral intensities of curves
described by Lorentzian and BWF (with the asymmetry parameter 1/q) functions normalized to the total intensity
of the band assigned to G band (S/Smax)
SWNT
SWNT:PSE
SWNT:PSE:GOX
ω,
cm−1
1521.7
1539.7
1551.2
1592.2
1601.4
∆Γ,
cm−1
30.1
12.6
15.0
15.4
31.9
−1/q
0.17
S/SG
×100%
3.1
5.7
11.1
48.8
31.3
ω,
cm−1
1520
1539.1
1550.0
1590.5
1598.5
∆Γ,
cm−1
29.2
12.2
14.6
15.1
31.4
−1/q
0.15
S/SG
×100%
3.9
6.1
10.3
46.5
33.2
ω,
cm−1
1518.4
1539.4
1550.5
1590.0
1598.0
∆Γ,
cm−1
37.8
16.5
13.2
15.4
28.9
−1/q
0.17
S/SG
×100%
5.3
8.3
7.4
45.8
33.2
Fig. 7.
scale) and PSE-GOX and SWNT (b, right scale)
Interaction energies between PSE and SWNT (a, left
and then equilibrated for 10 ns with a 1-fs step. After
modeling the total energy of interaction between SWNT
and the whole PSE:GOX complex, the interaction en-
ergies of SWNT and various parts of this complex were
determined. They are shown in Fig. 7. The equilibrated
structure of the SWNT:PSE:GOX complex is presented
in Fig. 8.
As is seen from Fig. 7, the energy of the interac-
tion between the pyrene fragment and the nanotube sur-
face in the SWNT:PSE:GOX complex (curve a) is in
the interval (−20 -- −25) kcal/mol. This result agrees
very well with the interaction energy between pyrene and
nanotubes calculated for the SWNT:PSE complex [14].
Moreover, Fig. 7 demonstrates that, during the whole
period of modeling, the pyrene fragment is strongly at-
tached to the nanotube. This evidences that one pyrene
"anchor" is able to hold efficiently the GOX enzyme on
the nanotube. The total interaction energy between a
nanotube and the PSE:GOX complex (Fig. 7, curve
b) (from −22 to −27 kcal/mol) is equal to the sum of
energies of interactions between PSE and a nanotube,
as well as between GOX and a nanotube. Thus, we
can determine the energy of interaction between GOX
Fig. 8. Equilibrated structure of the SWNT-PSE-GOX complex.
Water molecules are not shown
and a nanotube, by subtracting the SWNT:PSE inter-
action energy from the total interaction energy (Fig.
7, curve b). Thus, the energy of interaction between
GOX and a nanotube changes during the modeling from
−5 to −2 kcal/mol. This energy is due to the interac-
tions between a nanotube and outward side residues of
a GOX molecule. The mutual orientation of GOX and
the nanotube surface changes slowly during the model-
ing. As a result, the GOX -- nanotube interaction energy
changes as weakly. It may be concluded that the model-
ing of the SWNT:PSE:GOX complex demonstrates the
absence of strong interactions between GOX and the
nanotube surface. This fact allows us to suppose that
the GOX activity is not changed under the formation of
the SWNT:PSE:GOX complex.
3.4. Conductivity of single-walled carbon
nanotube networks: effects of environment
We carried out studies of the conductive properties of
the single-walled carbon nanotube network sprayed onto
a quartz substrate from their solution in dichloroben-
706
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
GLUCOSE OXIDASE IMMOBILIZATION ONTO CARBON NANOTUBE NETWORKING
Fig. 9. Current versus voltage for an SWNT network placed be-
tween two gold contacts separated by 10 µm gap (curve 1). Curve
2 corresponds to the conductivity of a sample after the adsorption
of PSE
zene. The dependence of the current through the car-
bon nanotube network in the dry state at variations of
the voltage (U ) between two contacts is shown in Fig.
9 (dotted line). The voltage is changed in ( -- 8 -- +8)
V range, the maximum current (I) runs up to 4 mA,
and the I(U ) dependence has nonlinear character. Most
likely, the nonlinearity is related to Schottky barriers,
which originate at the contact between nanotubes and a
gold contact or between nanotubes of different conduc-
tivities [22].
In addition, the I(U ) dependences man-
ifest a small hysteresis. To avoid the effects of a so-
lution on the volt-ampere characteristics, the presented
dependences were obtained in 4 -- 6 days after the fabri-
cation of a nanotube network or after the deposition of
biomolecules.
In this study, the effects of bioorganic compounds
(PSE and GOX) deposited on the carbon nanotube net-
work on its conductivity have been investigated. Af-
ter the deposition of the molecular interface (PSE) from
methanol and the drying of the film, its conductivity was
about 20% less than the initial value (Fig. 9, solid line).
Most probably, such a decrease is caused by adsorbed
PSE molecules, which induce the appearance of scatter-
ing centers for charge carriers on the nanotube surface,
as it was observed earlier [2]. It should be noted that
the following GOX adsorption (Fig. 9, inside curve 2)
has practically no effect on the conductivity of the nan-
otube network. It is obvious that this molecular inter-
face (PSE) isolates the enzyme from the nanotube sur-
face rather efficiently. The appearance of the hysteresis
for the I(U ) dependence in similar measurements was
explained earlier by the presence of charge traps on the
SiO2 surface. These traps are occupied under the pas-
Fig. 10. Time dependence of the maximal current (obtained at
8 V) for a SWNT network placed between two gold contacts sepa-
rated by 10 µm gap (black column). Dashed and shaded columns
correspond to the current through the network measured after the
deposition of PSE and the GOX immobilization, respectively
sage of a current in one direction and are not depleted
yet, when the current has the opposite direction [19].
In addition, the time dependence was built for the cur-
rent passing through the nanotube network, beginning
from the nanotube network fabrication up to the PSE de-
position and the immobilization of GOX (Fig. 10). As
is seen from Fig. 10, some spread of current values is ob-
served right away after the nanotube network treatment
with a bioorganic substance, which is caused by drying
the network after its wetting with methanol or water.
After some days, this spread of values becomes narrower.
A change in the conductivity is especially noticeable af-
ter the treatment of the nanotube network with water.
At once after the water evaporation, the film conductiv-
ity rises appreciably. Then it reduces in the course of
the time, but this takes a rather long time (some days
at room temperature).
4. Conclusion
The efficient immobilization of GOX onto a carbon nan-
otube network through the molecular interface formed
by PSE is carried out. This conclusion is based on the
analysis of AFM images of the network with the ad-
sorbed enzyme, whose globules locate mainly along a
nanotube.
The band corresponding to the high-frequency com-
ponent of the G mode in the RR spectrum of the nan-
otube with adsorbed PSE is downshifted by 0.7 cm−1
relative to this band in the spectrum of pristine nan-
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
707
V.A. KARACHEVTSEV, A.YU. GLAMAZDA, E.S. ZARUDNEV et al.
otubes. The analysis of the intensities of bands assigned
to the RBM of nanotubes with adsorbed PSE in compar-
ison with the spectrum of pristine SWNTs revealed the
intensity transformation, which can be explained by a
change of the resonance condition with variation of the
laser energy. Thus, we concluded that PSE molecules
create nanohybrids with SWNTs, which ensures the fur-
ther enzyme immobilization. As the RR spectrum of an
SWNT:PSE:GOX film does not essentially differ from
SWNT:PSE ones, this indicates that the molecular inter-
face (PSE) isolates the enzyme from nanotubes strongly
enough.
Our studies on the conductive properties of a single-
walled carbon nanotube network sprayed onto a quartz
substrate from a solution of nanotubes in dichloroben-
zene demonstrated that the I(U ) dependence has non-
linear character. Most likely, the nonlinearity is related
to Schottky barriers, which originate on the contact be-
tween nanotubes and the gold electrode, as well as be-
tween nanotubes with different conductivities.
The deposition of bioorganic compounds (PSE and
GOX) on the carbon nanotube network is accompanied
by a decrease of their conductivity. Most probably, such
a decrease is caused by adsorbed PSE molecules, which
induce the appearance of scattering centers for charge
carriers on the nanotube surface. The following GOX
adsorption has practically no effect on the conductiv-
ity of the nanotube network that evidences the reliable
isolation of the nanotube surface from the enzyme by
means of the molecular interface (PSE). While studying
the properties of carbon nanotube networks, whose sur-
face underwent the treatment with a solution, it is neces-
sary to consider the gradual desorption of a solvent from
the nanotube surface (especially, this concerns water).
The authors acknowledge the Computational Center
at B. Verkin Institute for Low Temperature Physics
and Engineering of the National Academy of Sciences
of Ukraine.
1. K. Balasubramanian and M. Burghard, Anal. Bioanal.
Chem. 385, 452 (2006).
7. S.S. Karajanagi, A.A. Vertegel, R.S. Kane,
and
J.S. Dordick, Langmuir 20, 11594 (2004).
8. R.J. Chen, Y. Zhang, D. Wang, and H. Dai, J. Am.
Chem. Soc. 123, 3838 (2001).
9. S.G. Stepanian, V.A. Karachevtsev, A.Yu. Glamazda,
U. Dettlaff-Weglikowska, and L. Adamowicz, Mol. Phys.
101, 2609 (2003).
10. W.E. Alvarez, F. Pompeo, J.E. Herrera, L. Balzano, and
D.E. Resasco, Chem. Mater. 14, 1853 (2002).
11. C.P. James, B. Rosemary, W. Wang, J. Gumbart,
E. Tajkhorshid, E. Villa, C. Chipot, R.D. Skeel, L. Kale,
and K. Schulten, J. Comput. Chem. 26, 1781 (2005).
12. A.D. MacKerell jr., D. Bashford, M. Bellott, R.L. Dun-
brack jr., J.D. Evanseck, M.J. Field, S. Fischer, J. Gao,
H. Guo, S. Ha, D. Joseph-McCarthy, L. Kuchnir,
K. Kuczera, F.T.K. Lau, C. Mattos, S. Michnick,
T. Ngo, D.T. Nguyen, B. Prodhom, W.E. Reiher,
B. Roux, M. Schlenkrich, J.C. Smith, R. Stote, J. Straub,
M. Watanabe, J. Wiorkiewicz-Kuczera, D. Yin, and
M. Karplus, J. Phys. Chem. B 102, 3586 (1998).
13. G. Wohlfahrt, S. Witt, J. Hendle, D. Schomburg,
H.M. Kalisz, and H.-J. Hecht, Acta Cryst. D 55, 969
(1999).
14. V.A. Karachevtsev, S.G. Stepanian, A.Yu. Glamazda,
M.V. Karachevtsev, V.V. Eremenko, O.S. Lytvyn, and
L. Adamowicz, J. Phys. Chem. C 115, 21072 (2011).
15. W. Humphrey, A. Dalke, and K. Schulten, J. Molec.
Graphics 14, 33 (1996).
16. S.G. Stepanian, M.V. Karachevtsev, A.Yu. Glamazda,
V.A. Karachevtsev, and L. Adamowicz, J. Phys. Chem.
A 113, 3621 (2009).
17. C. Fantini, A. Jorio, A.P. Santos, V.S.T. Peressinotto,
and M.A. Pimenta, Chem. Phys. Lett. 439, 138 (2007).
18. S.K. Doorn, J. Nanosci. Nanotech. 5, 1023 (2005).
19. S.G. Chou, H.B. Ribeiro, E.B. Barros, A.P. Santos,
D. Nezich, Ge.G. Samsonidze, C. Fantini, M.A. Pimenta,
A. Jorio, F. Plentz Filho, M.S. Dresselhaus, G. Dres-
selhaus, R. Saito, M. Zheng, G.B. Onoa, E.D. Semke,
A.K. Swan, M.S. Unlu, and B.B. Goldberg, Chem. Phys.
Lett. 397, 296 (2004).
20. M.S. Dresselhaus and P.C. Eklund, Adv. Phys. 49, 705
2. G. Gruner, Anal. Bioanal. Chem. 384, 322 (2006).
(2000).
3. S. Roy and Z. Gao, Nano Today 4, 318 (2009).
21. A.M. Rao, P.C. Eklund, S. Bandow, A. Thess, and
4. K. Besteman, J.O. Lee, F.G. Wiertz, H.A. Heering, and
R.E. Smalley, Nature 388, 257 (1997).
C. Dekker, Nano Lett. 3, 727 (2003).
5. X. Dong, C.M. Lau, A. Lohani, S.G. Mhaisalkar,
J. Kasim, Z.Shen, X. Ho, J.A. Rogers, and L-J. Li, Adv.
Mater. 20, 2389 (2008).
22. S. Heinze, J. Tersoff, R. Martel, V. Derycke, J. Ap-
penzeller, and P. Avouris, Phys. Rev. Lett. 89, 106801
(2002).
23. W. Kim, A. Javey, O. Vermesh, Q. Wang, and H. Dai,
6. P.W. Barone, S. Baik, D.A. Heller, and M.S. Strano, Na-
Nano Lett. 3, 193 (2003).
ture Mater. 4, 86 (2005).
Received 12.12.11
708
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
GLUCOSE OXIDASE IMMOBILIZATION ONTO CARBON NANOTUBE NETWORKING
IММОБIЛIЗАЦIЯ ГЛЮКОЗООКСИДАЗИ НА СIТКУ
ОДНОСТIННИХ ВУГЛЕЦЕВИХ НАНОТРУБОК
В.О. Карачевцев, О.Ю. Гламазда, Є.С. Заруднєв,
М.В. Карачевцев, В.С. Леонтьєв, О.С. Лiнник,
О.С. Литвин, О.М. Плохотниченко, С.Г. Степаньян
Р е з ю м е
При створеннi бiологiчних сенсорiв з використанням одностiн-
них вуглецевих нанотрубок (ОВНТ) треба вирiшити таку ва-
жливу проблему, як iммобiлiзацiя молекули, яка повинна роз-
пiзнати мiшень, на поверхнi нанотрубок. В данiй роботi про-
ведена iммобiлiзацiя ферменту глюкозооксидаза (ГОК) на по-
верхню сiтки нанотрубок, яка була одержана шляхом осадже-
ння нанотрубок з їх розчину у дiхлорбензолi за допомогою
спрей-методу. У ролi молекулярного iнтерфейсу було застосо-
вано сукцинiмiдний ефiр 1-пiренбутанової кислоти (ПСЕ), бi-
функцiональна молекула якого забезпечує хiмiчний зв'язок з
оболонкою ферменту, а друга її частина (пiренова) адсорбує-
ться на поверхню нанотрубки. Використання такого молеку-
лярного iнтерфейсу виключає, з одного боку, пряму адсорбцiю
ферменту на поверхню нанотрубки, яка знижує його актив-
нiсть, а з другого, забезпечує локалiзацiю ферменту поблизу
нанотрубки. Порiвняння спектрiв резонансного комбiнацiйно-
го розсiювання свiтла (РКРС) нанотрубок з їх спектром в ото-
ченнi ПСЕ вказує на створення наногiбриду молекулою ПСЕ
з нанотрубкою, що дає пiдставу для подальшої iммобiлiзацiї
ферментiв. Оскiльки спектри РКРС плiвок ОВНТ:ПСЕ:ГОК
суттєво не вiдрiзняються вiд спектрiв ОВНТ:ПСЕ, то можна
стверджувати, що молекулярний iнтерфейс ПСЕ достатньо мi-
цно iзолює фермент вiд нанотрубки. Ефективна iммобiлiзацiя
ферменту ГОК поблизу вуглецевої нанотрубки завдяки ПСЕ
пiдтверджується за допомогою зображень, отриманих атом-
силовим мiкроскопом. Молекулярна динамiка дозволила вста-
новити структури отриманих нанобiогiбридiв та енергiї мiж-
молекулярної взаємодiї мiж компонентами потрiйного компле-
ксу у водному оточеннi. Було також дослiджено провiднi вла-
стивостi сiтки ОВНТ з адсорбованими молекулами ПСЕ та
ГОК. Адсорбцiя молекул ПСЕ на сiтку з ОВНТ супроводжу-
ється зменшенням провiдностi, яке, скорiш за все, пов'язано
з появою розсiювальних центрiв для носiїв заряду у нанотруб-
ках.
ISSN 2071-0186. Ukr. J. Phys. 2012. Vol. 57, No. 7
709
|
1803.03425 | 2 | 1803 | 2018-11-09T16:17:08 | Role of hydrodynamic flows in chemically driven droplet division | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.CB"
] | We study the hydrodynamics and shape changes of chemically active droplets. In non-spherical droplets, surface tension generates hydrodynamic flows that drive liquid droplets into a spherical shape. Here we show that spherical droplets that are maintained away from thermodynamic equilibrium by chemical reactions may not remain spherical but can undergo a shape instability which can lead to spontaneous droplet division. In this case chemical activity acts against surface tension and tension-induced hydrodynamic flows. By combining low Reynolds-number hydrodynamics with phase separation dynamics and chemical reaction kinetics we determine stability diagrams of spherical droplets as a function of dimensionless viscosity and reaction parameters. We determine concentration and flow fields inside and outside the droplets during shape changes and division. Our work shows that hydrodynamic flows tends to stabilize spherical shapes but that droplet division occurs for sufficiently strong chemical driving, sufficiently large droplet viscosity or sufficiently small surface tension. Active droplets could provide simple models for prebiotic protocells that are able to proliferate. Our work captures the key hydrodynamics of droplet division that could be observable in chemically active colloidal droplets. | physics.bio-ph | physics |
Role of hydrodynamic flows in chemically driven droplet division
Rabea Seyboldt1, ∗ and Frank Julicher1, 2, †
1Max Planck Institute for the Physics of Complex Systems, 01187 Dresden, Germany
2Center for Systems Biology Dresden, 01307 Dresden, Germany
We study the hydrodynamics and shape changes of chemically active droplets. In non-spherical
droplets, surface tension generates hydrodynamic flows that drive liquid droplets into a spherical
shape. Here we show that spherical droplets that are maintained away from thermodynamic equi-
librium by chemical reactions may not remain spherical but can undergo a shape instability which
can lead to spontaneous droplet division. In this case chemical activity acts against surface ten-
sion and tension-induced hydrodynamic flows. By combining low Reynolds-number hydrodynamics
with phase separation dynamics and chemical reaction kinetics we determine stability diagrams of
spherical droplets as a function of dimensionless viscosity and reaction parameters. We determine
concentration and flow fields inside and outside the droplets during shape changes and division. Our
work shows that hydrodynamic flows tends to stabilize spherical shapes but that droplet division
occurs for sufficiently strong chemical driving, sufficiently large droplet viscosity or sufficiently small
surface tension. Active droplets could provide simple models for prebiotic protocells that are able to
proliferate. Our work captures the key hydrodynamics of droplet division that could be observable
in chemically active colloidal droplets.
Keywords: hydrodynamic flows, phase separation, nonequilibrium shape dynamics
Living cells are compartmentalized in order to organize
their biochemistry in space. Many cellular compartments
do not possess membranes and are formed by the assem-
bly of proteins and RNA in compact condensates1 -- 16.
Such condensates often have liquid like properties and
resemble droplets that form by phase separation of a
complex mixture1,11 -- 13. Indeed protein droplets are ob-
served to form in vitro by phase separation in physiologi-
cal buffer13,15,17 -- 19. Such droplets can organize chemical
reactions in space, and the droplet dynamics can in turn
be influenced by the reactions, as has been shown both in
theory8,20 -- 26 and experiments13,15,17 -- 19,27,28. The ubiq-
uitous nature of RNA-protein condensates in a large va-
riety of different cells and organisms suggests that the
physical chemistry of macromolecular phase separation
represents an evolutionary old mechanism for the com-
partmentalization of chemistry and that droplet forma-
tion could have played a key role at the origins of life and
the emergence of prebiotic protocells15,18,29 -- 40.
A minimal model of a protocell consists of a droplet
that turns over by a chemical reaction and is constantly
supplied with droplet material by diffusion from the
outside39.
In such a scenario droplets are maintained
away from thermodynamic equilibrium and can reach a
non-equilibrium steady state with a radius that is set
by reaction parameters26. An interesting possibility is
that the spherical shape of active droplets becomes un-
stable and droplets spontaneously divide in two smaller
daughters drops, providing a physical mechanism for the
division of prebiotic cells39. Such droplet dynamics is a
hydrodynamic problem because surface tension in non-
spherical droplets drives hydrodynamic flows that redis-
∗ Email: [email protected]
† Email: [email protected]
FIG. 1.
Chemically active droplet described by an effec-
tive droplet model. Shown is the concentration field c (blue
and green color) of a stationary droplet (interface in black).
Chemical reactions B → A create a sink of droplet material B
in the droplet, and reactions A → B create a supersaturation
of droplet material in the A-rich phase outside. This creates
concentration gradients of B, which drive diffusion fluxes of
droplet material, while A flows in the opposite direction. The
stationary droplet size results from the balance of the fluxes
(Parameters: = 0.176, A = 10−2,
across the interface.
η+/η− = 1, k+/k− = 1, ν−/(k−∆c) = 1, D+/D− = 1,
β− = β+, c(0)
+ = 0)
tribute material and deform the droplet shape41 -- 44.
Here we develop a hydrodynamic theory of the dynam-
ics of chemically active droplets. We show that chemical
reactions in active droplets can perform work against sur-
face tension and flows, giving rise to a shape instability
that can result in droplet division. We investigate the
conditions for which droplets divide and determine hy-
drodynamic flow fields of dividing droplets.
We consider an incompressible liquid consisting of
B-richA-richdroplet material B and solvent component A which can
phase separate. The local composition is characterized
by the concentration field c(x) of component B. Volume
preserving chemical reactions can transform component
A into component B and back, A (cid:10) B. For simplicity, we
first discuss an effective droplet model. A single droplet
characterized by high concentration c of component B co-
exists with the surrounding fluid that mainly consists of
A and contains B at low concentration, see Fig. 1. Both
phases are separated by a sharp interface. The concentra-
tion of B satisfies a balance equation, where the chemical
reaction provide a source or sink term s±(c),
∂tc + ∇ · j = s±(c)
j = −D±∇c + vc .
(1)
(2)
Here, the indices + and − refer to quantities outside and
inside the droplet, respectively. The flux j consists of
advection by the fluid velocity v and a diffusion flux,
where D± denotes the diffusion constant of the droplet
material in the two phases.
is
by
The
reaction
chemical
described
the
rate s±(c) which in gen-
concentration-dependent
eral
is a nonlinear function of c. For simplicity, we
linearize the chemical reaction rates in the vicinity of
reference concentrations c(0)± in each phase:
s±(c) (cid:39) −k±(c − c(0)± ) ± ν± ,
(3)
where c(0)± are the equilibrium concentrations that coex-
ist at equilibrium across a flat interface. We have defined
the reaction rate ν± = s(c(0)± ) and the reaction constants
k± = ds(c(0)± )/dc. We focus on the case of positive coef-
ficients k± > 0 and ν± > 0, where B is produced outside
the droplet, and degraded inside, see Fig. 1.
The hydrodynamic flow velocity v obeys Stokes equa-
tion of an incompressible fluid,
η±∇2v = ∇p ,
(4)
which accounts for momentum conservation ∂ασαβ = 0,
where the stress tensor is given by σαβ = η±(∂αvβ +
∂βvα)− pδαβ. Here η± denotes the fluid shear viscosities
inside and outside of the droplet. The pressure p plays
the role of a Lagrange multiplier to impose the constraint
∇ · v = 0.
The bulk equations (1-4) are connected by boundary
conditions at the droplet interface which we parameterize
in spherical coordinates by the radial interface position
R(θ, φ) as a function of the polar and azimuthal angles θ
and φ. The stress boundary conditions read
nn(R) − σ−
σ+
nt(R) − σ−
σ+
nn(R) = 2γH(R)
nt(R) = 0 ,
(5)
(6)
where H(R) is the local mean curvature of the interface
and γ is the droplet surface tension. The stresses at the
interface on the inner and outer side of the droplet are de-
noted by σ±
αβ(R). The tensor indices n and t refer to ten-
sor components normal and tangential to the interface,
2
nn = nασ±
αβnβ and σ±
respectively. The normal and tangential tensor compo-
nents are defined as σ±
αβtβ,
where nα is a unit vector normal to the surface and tα is
a unit vector tangent to the surface. Eq (6) is valid for all
tangent vectors and summation over repeated indices is
implied. Using no-slip boundary conditions, the velocity
field is continuous at the interface,
nt = nασ±
v+(R) = v−(R) .
(7)
The concentration field c is discontinuous across the in-
terface with values given by
c−(R) = c(0)− + β−γH(R)
c+(R) = c(0)
+ + β+γH(R)
(8)
(9)
which are set by the physics of phase coexistence and a lo-
cal equilibrium assumption. The coefficients β± describe
the effects of the Laplace pressure on the equilibrium con-
centrations at phase coexistence. In the presence of fluxes
at the interface, the interface moves in normal direction.
The radial growth velocity is
(R) − j+(R)
−
· j
c−(R) − c+(R)
n
n · er
dR
dt
(10)
=
,
where n is a unit vector normal to the surface and er is
a unit vector in radial direction. Eq. (10) captures both
convection of the interface by flows and droplet growth
and shrinkage by addition or removal of material.
We find nonequilibrium steady state solutions to Equa-
tions (1-10) with a spherical droplet of stationary radius
¯R and stationary concentration field ¯c(r), where r is the
radial coordinate, see Appendix A. The stationary pres-
sure ¯p exhibits a jump 2γ/ ¯R across the interface and no
hydrodynamic flows exist, ¯v = 0. An example for a sta-
ble non-equilibrium steady state with steady state con-
centration profile inside and outside the droplet of radius
¯R is shown in Fig. 1.
We first discuss the properties of these stationary
states as a function of external supersaturation =
ν+/(k+∆c) and the dimensionless reaction rate A =
ν−τ /∆c inside the droplet. The supersaturation is in
our system generated by reactions outside the droplet
and in steady state corresponds to the concentration for
which s+ = 0. Here, ∆c = c(0)− − c(0)
+ and we have intro-
duced the time scale τ = w2/D+, where w = 6β+γ/∆c
is a characteristic length scale. The stationary radii as
a function of supersaturation are shown In Fig. 2A-C
as solid lines for different values of A. For values of
smaller than a threshold value 0, no stationary radius
exists. For values > 0 two steady state radii ¯Rc and
¯Rs exist, which become equal at 0 where they approach
a value ¯R0. The smaller steady state radius ¯Rc is a crit-
ical nucleation radius similar to the critical droplet radii
found in passive systems. The larger radius denoted ¯Rs
stems from the interplay of phase separation and chemi-
cal reactions26,39. As the supersaturation reaches a value
∞ =(cid:112)(D−k−)/(D+k+)ν−/(k−∆c), the stationary ra-
dius ¯Rs diverges.
3
FIG. 2. Stationary radii and onset of shape instability.
A -- C: Stationary radius as a function of supersaturation for different reaction amplitudes A = 10−8, 10−7, . . . , 101. The station-
ary radii (lines) are independent of the dimensionless viscosity F = wη−/(γτ ), while the onset of instability (red dots, connected
by dotted red line) for the different curves varies in the three figures, which show dimensionless viscosities F = ∞, 1000, 10 (left
to right). The blue line colors mark stable, the red ones unstable stationary radii with respect to the elongational l = 2 mode.
In panel B the scaling behavior of the nucleation radius ¯Rc and the stationary radius ¯Rs are indicated.
D -- F: Stability diagram of stationary droplets of size ¯Rs, as a function of reaction amplitude A and supersaturation for
different dimensionless viscosities F = ∞, 1000, 10 (left to right). For small supersaturation and large reaction amplitudes, no
stationary radius exists (white). For large supersaturation, the stationary radius diverges (gray). In the region between these
regimes, the stationary solution can be stable (blue) or unstable (red) with respect to shape perturbations of the l = 2 mode.
For decreasing F , the stable regime grows, and the minimal supersaturation ∗ at which an instability can be found increases,
as well as the corresponding reaction amplitude A∗. The scaling relations (dashed lines) for the regime of stable droplets and
the onset of instability are indicated, with prefactors according to A 5. (Parameters: η+/η− = 1, k+/k− = 1, ν−/(k−∆c) = 1,
D+/D− = 1, β− = β+, c(0)
+ = 0)
We can find simple expressions for the stationary
radii in the limit of small A while keeping the ratios
ν−/(k−∆c) and k+/k− of reaction parameters fixed. In
this limit, the chemical reactions fluxes vanish as s± ∝ A
and the threshold value 0 vanishes as 0 ∝ A1/3. The
critical nucleation radius then behaves as ¯Rc (cid:39) w/(6)
and the larger steady state radius ¯Rs (cid:39) w(3A)1/2 where
0 (cid:28) (cid:28) ∞, see Fig. 2B and A 5.
The steady state solutions are independent on the fluid
viscosity, however the droplet dynamics is affected by hy-
drodynamic effects. We now investigate the role of hy-
drodynamic flows on chemically driven shape instabilities
that can give rise to droplet division. We perform a lin-
ear stability analysis at the stationary state given by ¯X =
(¯c, ¯R, ¯p, ¯v) for small perturbations δX = (δc, δR, δp, δv).
The dynamics of these perturbations can be represented
using eigenmodes
(cid:88)
δX =
nlmXnlmeµnlmt ,
(11)
n,l,m
with Xnlm = (cnlYlm, ¯RYlm, plYlm, vlm), where Ylm(θ, φ)
are spherical harmonics with angular mode indices with
SupersaturationSupersaturationSupersaturationSupersaturationSupersaturationSupersaturationStationary radiusStationary radiusStationary radiusDimensionless reaction rateDimensionless reaction rateDimensionless reaction rateABCDEFno stationarydropletsshape instabilitystable dropletsno stationarydropletsno stationarydropletsstable dropletsstable dropletsshape instabilityshape instabilityl = 0, 1, . . . and m = −l, . . . , l. The index n = 0, 1, . . .
denotes radial modes. The eigenmodes exhibit an expo-
nential time dependence with a relaxation rate given by
the eigenvalue µnlm. The mode amplitudes are denoted
nlm. The concentration modes are characterized by the
radial functions cnl(r). The pressure modes are described
by pl(r) and the velocity modes vlm(r, θ, ϕ) can be ex-
pressed as
lm Φlm .
vlm = vr
lmY lm + v(1)
lm Ψlm + v(2)
lm(r), v(1)
lm (r) and v(2)
(12)
where Y lm(θ, ϕ) = erYlm, Ψlm(θ, ϕ) = r∇Ylm and
Φlm(θ, ϕ) = er × Ψlm are vector spherical harmonics45
and the radial functions vr
lm (r) char-
acterize the velocity field. The radial functions can be
obtained by solving the linearized dynamic equations us-
ing the corresponding boundary conditions, see A. The
Stokes equation can be solved for a given shape pertur-
bation independent of the concentration field so that the
velocity field and pressure field is independent of the ra-
dial mode n. The radial part of the concentration field
obeys a Helmholtz equation with an inhomogeneity that
stems from hydrodynamic flows. The homogeneous part
is solved by modified spherical Bessel functions and the
inhomogeneous solution can be found using Greens func-
tions. Using the dynamic equation for the shape changes
of the droplet Eq. (10), we obtain an equation for the
eigenvalue µnlm,
(cid:18)
µnlm =
l ( ¯R)
vr
¯R
(cid:18)
+
D+
∆c
− D−
∆c
c(cid:48)
nl( ¯R+)
¯c(cid:48)(cid:48)( ¯R+) +
¯R
c(cid:48)
nl( ¯R−)
¯c(cid:48)(cid:48)( ¯R−) +
(cid:19)
(cid:19)
.
¯R
(13)
Here, the primes denote radial derivatives. Note that
Eq. (13) is an implicit equation for the eigenvalues µnlm
because the radial concentration modes cnl(r) depend
on µnlm, see A. Eq. (13) is independent of the index m,
therefore the degeneracy of an eigenvalue µnl is at least
2l + 1. When all µnl are negative, the spherical shape
is stable. The modes with l = 0 correspond to changes
in droplet size without flows. They are always stable for
¯R = ¯Rs and unstable for ¯R = ¯Rc. Thus droplet smaller
than ¯Rc will vanish and droplets larger will grow towards
the size ¯Rs. Thus we consider the stability of ¯R = ¯Rs
in the following. The modes with l = 1 do not involve
shape deformations of the droplet and are thus not asso-
ciated with flows. There always exists a marginal mode
with µl=1 = 0 corresponding to overall translations where
the droplet and all concentration fields are displaced and
then stay in the new position. Here we consider shape in-
stabilities for which a mode with l > 1 becomes unstable.
Because shape deformations induce flows, this instability
depends on the dimensionless viscosity F = wη−/(γτ ), as
well as the ratio of viscosities in the two phases, η+/η−.
If we increase the supersaturation while keeping the
other parameters fixed, the steady state can become un-
stable with respect to the mode l = 2 for a critical value
4
= c. In Fig. 2A-C, the onset of instability µ = 0 for
the largest eigenvalue µ of the stationary radius is shown
as a red dot, and unstable radii are indicated by red
lines. Different lines correspond to different supersatu-
rations, and the panels show different values of F .
In
Fig. 2D-E, the corresponding stability diagrams of sta-
tionary droplets are shown as a function of the supersat-
uration and the reaction amplitude for different values of
F . For large A and small , no stationary radius exists
(white regions), so that any droplet would shrink and dis-
appear. For large , the stationary state diverges (gray
regions). Spherical droplets are stable in the blue regions.
Stationary spherical droplets are unstable inside the red
region, the surrounding black line marks the shape insta-
bility with respect to the l=2 mode. The region where
spherical droplets undergo a shape instability exists for
≥ ∗, which depends on F . The value of A for which
the shape instability occurs at = ∗ is denoted A∗, see
Fig 2E.
For small A, the onset of instability can be describes
by simple scaling behaviors. In the limit of small A and
for (cid:28) ∞, we find ∗ ∝ F −1/2 and A = A∗ with A∗ ∼
F −3/2 (compare Fig. 2E-F). For A < A∗, hydrodynamic
flows govern the onset of instability which occurs at a
value of A which behaves as A ∝ −1F −2. For A >
A∗, hydrodynamic flows can be neglected as compared
to diffusion fluxes and the onset of instability occurs for
A ∝ 3. These two scaling regimes are indicated in in
Fig. 2D-F by dashed lines. A derivation of these results
including prefactors is given in A 5.
We next address the question whether the shape in-
stability found in the linear stability analysis can indeed
give rise to droplet divisions in the presence of hydro-
dynamic flows in the nonlinear regime of the dynamics.
We use a Cahn-Hilliard model46 for phase separation dy-
namics, extended to include chemical reactions and hy-
drodynamic flows, that can capture topological changes
of the interface. We include chemical reactions via a
source term linear in the concentration as well as advec-
tion by the hydrodynamic flow which is described by the
incompressible Stokes equation. Using a semi-spectral
method47, we obtain numerical solutions in a cubic box
with no-flux boundary conditions, see B.
Starting from a weakly deformed spherical droplet, we
find regimes where the droplet disappears, where it re-
laxes to a stable spherical shape and where it undergoes
a shape instability, consistent with the linear stability
analysis of the effective droplet model. The transitions
between these regimes occur for parameter values close
to those predicted by the linear stability analysis. In the
unstable regime, droplets typically divide. This shows
that the droplet division reported previously can also oc-
curs in the presence of hydrodynamic flows. Fig. 3 shows
snapshots of the droplet shape together with correspond-
ing hydrodynamic flow fields on the symmetry plane of a
dividing droplet at different times. At early times when
the droplet deformation is weak, the flow field is similar
to the l = 2 mode obtained from the linear theory, Fig. 3
5
between the daughter droplets has very small magnitude,
while strong flows at the outer sides move the droplets
apart Fig. 3 D.
This example shows that division of active droplets
can occur even if hydrodynamic flows that oppose divi-
sion are taken into account. Because flows act in oppo-
sition to the initial deformation of the sphere, the linear
stability analysis already provides the key information of
whether droplet division can occur for a given value of di-
mensionless viscosity F , see Fig. 2. This raises the ques-
tion under what experimental conditions active droplets
would become unstable and division could be observed.
Ignoring hydrodynamic flows, F → ∞, it was shown that
oil-water droplets and soft colloidal liquids or p-granules
with sizes of a few micrometers could divide in the pres-
ence of chemical reactions39. To address the influence
of hydrodynamic flows, we have to estimate the dimen-
sionless viscosity F = wη−/(γτ ) (cid:39) kBT /(6πγwa), where
we have used τ = w2/D and D (cid:39) kBT /(6πηa) with
molecular radius a. Thus, F is an equilibrium property
of the phase separating fluid. For an oil-water system,
we estimate F ≈ 0.1, see C. For soft colloidal liquids or
p-granules, we estimate values between F ≈ 10 − 104.
We can discuss these values using the stability diagrams
in Fig. 2D-F. Oil-water like droplets with F ≈ 0.1 are
unlikely to divide, as the unstable region in the stability
diagram is very narrow. For soft colloidal systems with
F ≈ 10 − 104, droplet division might be experimentally
observable. We can estimate typical reaction rates re-
quired for division to occur based on the reaction rate A∗
for which the range of supersaturation is maximal. The
value of A∗ corresponds to a reaction rate in the droplet
of the order of ν− = 10−4mM/s, see C. A comparison
with reported enzymatic reaction rates48 suggests that
such values can be achieved in real systems.
We have shown that spontaneous division of chemi-
cally active droplets involves mechanical work against
surface tension as droplets deform. Active droplets thus
can transduce chemical energy to mechanical work and
droplet division is therefore a mechano-chemical process.
The surface tension of the droplet creates pressure gra-
dients as the droplet becomes non-spherical that lead to
hydrodynamic flows. Because the flows generated act
against the shape deformation, droplets divide only for
sufficiently large viscosity or sufficiently small surface
tension and sufficiently large reaction rates. We show
that the dependence of the onset of stability on parame-
ters is captured for small reaction fluxes by simple scaling
relations. Our work shows that droplet division would
be suppressed in oil-water systems due to large surface
tension and low viscosity. However it could be realized
in soft colloidal systems for chemical reaction parameters
that could be achieved experimentally. Furthermore flux-
driven droplet divisions could be observable in biological
systems, as both chemical reactions and phase-separating
membrane-less organelles with low surface tensions can
be found within cells.
FIG. 3.
Numerical solution in 3d of an extended Cahn-
Hilliard model with chemical reactions and hydrodynamic
flows reveals that droplets can divide despite the presence of
hydrodynamic flows. Panels A-D correspond to time points
t/τ = 100, 2100, 2700, 2800, respectively, where τ = w2/D
is a diffusion time, with diffusion constant D and interfacial
width w. The dynamic equations were solved numerically in
a three-dimensional box. Shown are two-dimensional cross-
sections of the droplet shape (black) together with streamlines
(grey). Arrows (colored) indicate the direction and magni-
tude of the flow (normalized by respective maximal veloci-
ties vmax · w/D = 0.0016(in A), 0.0048(B), 0.0034(C) and
0.0047(D) ). (Parameters: F = 24, A = 8 · 10−3, = 0.2,
η−/η+ = 1, c(0)
+ /∆c = 0, k+/k− = 1, ν−/(k−∆c) = 0.8)
A. As the droplet elongates and its waistline shrinks, the
flow field becomes more complex, see Fig. 3 B,C. The
flow field shown in Fig. 3 C exhibits two additional vortex
lines that form rings around the axis of rotational sym-
metry. Similarly, after division, two further vortex rings
occur, see Fig. 3 D. Interestingly, for small deformations
the hydrodynamic flow direction opposes the directions of
interface motion at the main droplet axes, see Fig. 3 A,B.
For larger deformations at later times the flow switches
its direction along the long droplet axis where it assists
interface motion. At the waistline, the flow velocity be-
comes small, see Fig. 3 C. After division, the flow field
ABCDAppendix A: Effective droplet model with hydrodynamic flows
1. Stationary state of a spherical active droplet
6
Here, we discuss stationary solutions to equations (1-10) in the main text with spherical symmetry and without
hydrodynamic flows ¯v = 0, where the bar indicates a steady state value. In this case, the pressure is constant both
inside and outside the droplet, with a pressure difference due to Laplace pressure between the inside and outside of
the droplets,
The steady state concentration profiles in the presence of chemical reactions are given by39,
¯p− = ¯p+ +
2γ
¯R
.
+ c(0)
+ + A+k0(r/l+)
ν+
¯c+(r) = +
k+
¯c−(r) = − ν−
k−
(A3)
where i0(x) = 2 sinh(x)/x and k0(x) = e−x/x denote modified spherical Bessel functions of order zero of the first
and second kind, respectively. The characteristic length scales l± = (D±/k±)1/2 are set by reaction rate constants
and diffusion coefficients. The parameters A± are determined by the boundary condition at the droplet interface, Eq.
(8-9) in the main text,
+ c(0)− + A−i0(r/l−) ,
A+ =
A− =
Stationarity of the droplet radius ¯R implies
(cid:18) γβ+
(cid:18) γβ−
¯R
¯R
(cid:19)
(cid:19)
− ν+
k+
ν−
k−
+
1
k0( ¯R/l+)
1
i0( ¯R/l+)
.
D+¯c(cid:48)
+( ¯R) = D−¯c(cid:48)
−( ¯R) ,
see Eq. (10) in the main text. Note that this equation typically has zero, one or two solutions for a given set of
parameters.
2. Linearized dynamics
We introduce small perturbations to the spherically symmetric stationary state, with p = ¯p + δp, v = δv, c = ¯c + δc
and R = ¯R + δR and write the dynamics of these perturbations to linear order. The linearized dynamics reads
(A1)
(A2)
(A4)
(A5)
(A6)
(A7)
(A8)
(A9)
(A10)
(A11)
∇δp = η±∆δv
∇ · δv = 0
∂tδc = −δvr¯c(cid:48) + D±∇2δc − k±δc
∂tδR = δvr( ¯R) +
−( ¯R)(cid:3) δR
(cid:2)D+∂rδc+( ¯R) − D−∂rδc−( ¯R)(cid:3) .
(cid:2)D+¯c(cid:48)(cid:48)
+( ¯R) − D−¯c(cid:48)(cid:48)
1
∆c
+
1
∆c
Here δvr denotes the radial part of the hydrodynamic velocity. With δc− and δc+ we denote perturbations of the
concentration field inside and outside the droplet. The same notation holds for the other fields. In this linear analysis,
boundary conditions apply at the stationary radius ¯R,
δc±( ¯R) = β±γδH − ¯c(cid:48)
±( ¯R)δR ,
The linearized dynamics can be decomposed in spherical harmonics, see Eq (11) in the main text. The curvature
with perturbation of the curvature δH = H(R) − H( ¯R).
perturbation then takes the form
δH =
with hl = (l2 + l − 2)/2.
nlm
(cid:88)
hl
¯R
nlmYlm ,
(A12)
3. Hydrodynamic eigenmodes of the linearized dynamics
7
We can expand the hydrodynamic eigenmodes using a basis of vector spherical harmonics, see Eq. (12) in the main
text. The velocity boundary conditions Eq. (7) in the main text for the mode amplitudes read
0 = vr+
0 = v(1)+
0 = v(2)+
lm ( ¯R) − vr−
lm ( ¯R)
lm ( ¯R) − v(1)−
lm ( ¯R)
lm ( ¯R) − v(2)−
lm ( ¯R) .
The stress boundary conditions (see Eq. (5-6) in the main text) at the interface read
lm( ¯R) − 2γlm
hl
¯R
0 = 2η+(vr+
0 = η+
(cid:34)
(cid:35)
lm )(cid:48)( ¯R) + p−
lm ( ¯R)
(cid:35)
lm )(cid:48)( ¯R) − p+
lm )(cid:48)( ¯R) +
(cid:34)
(v(1)+
lm( ¯R) − 2η−(vr−
− v(1)+
lm ( ¯R)
vr+
¯R
¯R
− v(1)−
vr−
lm ( ¯R)
lm ( ¯R)
(cid:35)
(cid:34)
¯R
¯R
lm )(cid:48)( ¯R) − v(2)+
(v(2)+
¯R
(v(1)−
lm )(cid:48)( ¯R) +
lm ( ¯R)
− η−
−η−
(cid:34)
0 = η+
lm )(cid:48)( ¯R) − v(2)−
(v(2)−
lm ( ¯R)
¯R
(cid:35)
.
(A13)
(A14)
(A15)
(A16)
(A17)
(A18)
(A19)
(A20)
(A21)
(A22)
(A23)
(A24)
(A25)
(A26)
(A27)
We solve the radial profiles of the modes with a polynomial ansatz and exclude functions that diverge for r → 0 or
r → ∞ inside and outside the droplet, respectively. The pressure is then given by
(cid:16) r
(cid:17)l+1
(cid:17)−l
(cid:16) r
p−
lm(r) = γfA
lm(r) = −γfB
p+
R
R
.
For the hydrodynamic flow velocity we obtain
(cid:20)
(cid:16) r
(cid:20) l + 3
fC1
R
l(l + 1)
(cid:17)l+1 − fC3
(cid:16) r
(cid:17)l−1(cid:21)
(cid:16) r
(cid:17)l+1 − l + 1
fC1
R
R
l(l + 1)
(cid:16) r
(cid:17)l−1(cid:21)
R
fC3
γ
η−
γ
η−
vr−
lm (r) =
v(1)−
lm (r) =
v(2)−
lm (r) = 0
and
vr+
lm (r) =
γ
η−
γ
η−
v(2)+
lm (r) = 0 .
v(1)+
lm (r) =
(cid:17)−l−2(cid:21)
(cid:20)
(cid:16) r
(cid:20) l − 2
−fC2
R
(cid:17)−l
(cid:16) r
fC2
l(l + 1)
R
(cid:16) r
(cid:17)−l − 1
R
+ fC4
fC4
l + 1
(cid:16) r
(cid:17)−l−2(cid:21)
R
Here, we have defined
fA =
fB =
fC1 =
fC2 =
fC3 =
fC4 =
(l − 1) (l + 1) (l + 2) (2l + 3)
∆ (2l2 + 4l) + (2l2 + 4l + 3)
l (l − 1) (l + 2) (2l − 1)
(2l2 + 1) + (2l2 − 2) /∆
1
2
1
2
∆ (2l2 + 4l) + (2l2 + 4l + 3)
l (l − 1) (l + 1) (l + 2)
∆ (2l2 + 1) + (2l2 − 2)
l (l − 1) (l + 1) (l + 2)
l (l − 1) (l + 1) (l + 2)(cid:0)∆(cid:0)2l2 + 4l + 3(cid:1) +(cid:0)2l2 + 4l(cid:1)(cid:1)
l (l − 1) (l + 1) (l + 2)(cid:0)∆(cid:0)2l2 − 2(cid:1) +(cid:0)2l2 + 1(cid:1)(cid:1)
(∆ (2l2 + 1) + (2l2 − 2)) (∆ (2l2 + 4l) + (2l2 + 4l + 3))
(∆ (2l2 + 1) + (2l2 − 2)) (∆ (2l2 + 4l) + (2l2 + 4l + 3))
1
2
1
2
where ∆ = η+/η− denotes the ratio of the viscosities inside and outside the droplet.
4. Concentration eigenmodes
The equation for the radial part of the concentration eigenmode is
l (r)¯c(cid:48)(r) =
vr
1
D±
d
dr
r2 d
dr
− λ±2
nl − l(l + 1)
r2
(cid:20) 1
r2
(cid:21)
cnl(r)
with
The boundary conditions at ¯R are
λ±2
nl = (k± + µnl)/D± .
cnl( ¯R±) = γβ±
− ¯R ¯c(cid:48)( ¯R±) .
hl
¯R
The left-hand side of Eq. (A34) constitutes an inhomogeneity
l (r) = − 1
f±
D±
l (r)¯c(cid:48)(r) .
vr
,
8
(A28)
(A29)
(A30)
(A31)
(A32)
(A33)
(A34)
(A35)
(A36)
(A37)
The solution c±
constructed from a particular solution c±
homogeneous equation with f±
as
nl(r) of the inhomogeneous equation (A34) that satisfies the boundary condition Eq. (A36) can be
nl,h(r) of the
l = 0 are added to satisfy the boundary conditions, Eq. (A36). This can be expressed
nl,p(r) of the inhomogeneous equation to which solutions c±
c−
nl(r) = α−
c+
nl(r) = α+
nlc−
nlc+
nl,h(r) + c−
nl,h(r) + c+
nl,p(r)
nl,p(r) ,
where the coefficients α± read
α±
nl =
nl,p( ¯R)
l − c±
a±
c±
nl,h( ¯R)
,
(A38)
(A39)
(A40)
nl( ¯R).
l = c±
with a±
equation (A34) for λ±2
equation which is solved by modified spherical Bessel functions, c−
We are especially interested in the case of unstable modes with µnl > 0. Therefore we focus on the solution of
l = 0 is a modified Helmholtz
nlr), where il
nl > 0 and k± > 0. In this case, the homogeneous equation with f±
nl,h(r) = il(λ−
nl,h(r) = kl(λ+
nlr) and c+
and kl denote the modified spherical Bessel functions of first and second order, respectively. The particular solution
of the inhomogeneous equation can be obtained by a Green's function approach,
c−
l,p(r) = λ−
nlkl(λ−
nlr)
nlr2)f−
l (r2)r2
2
+λ−
nlil(λ−
nlr)
nlr2)f−
l (r2)r2
c+
l,p(r) = λ+
nlkl(λ+
nlr)
nlr2)f +
l (r2)r2
2
+λ+
nlil(λ+
nlr)
nlr2)f +
l (r2)r2
2
0
(cid:90) r
(cid:2)il(λ−
(cid:90) ¯R
(cid:2)kl(λ−
(cid:90) r
(cid:2)il(λ+
(cid:90) ∞
(cid:2)kl(λ+
¯R
r
r
(cid:3) dr2
(cid:3) dr2
(cid:3) dr2
(cid:3) dr2 ,
2
9
(A41)
(A42)
(A43)
(A44)
(A45)
(A46)
with the radial part of the inhomogeneity f±
l (r) given by Eq. (A37). The explicit calculation of these functions has
to be handled with care, since the functions kl and il have divergences for large and small arguments r that cancel in
the final result but can still lead to numerical difficulties when evaluated directly.
The derivative of the concentration profile at ¯R can be expressed as
−
l
c(cid:48)
nl( ¯R−) = a
c(cid:48)
nl( ¯R+) = a+
¯R gl,i(λ−
¯R gl,k(λ+
nl
l
¯R) +
nl
¯R) +
c
−
l,p( ¯R)
¯R
l,p( ¯R)
c+
¯R
·(cid:2)gl,k(λ−
·(cid:2)gl,i(λ+
nl
nl
¯R) − gl,i(λ−
¯R) − gl,k(λ+
nl
nl
¯R)(cid:3)
¯R)(cid:3) ,
with
gl,i(x) = xi(cid:48)
gl,k(x) = xk(cid:48)
l(x)
il(x)
l(x)
kl(x) .
Using the equation for the shape perturbations (A10), and using Eqns (A43) and (A44), we obtain Eq. (13) in the
main text. This equation determines the eigenvalue µnlm of the hydrodynamic modes.
5. Scaling relations in the limit of small reaction fluxes
In the limit of small chemical reaction fluxes s± we obtain simple scaling expressions for stationary radii and their
shape instability conditions. Here we present the method and discuss the results.
a. Stationary radius
Here we discuss the stationary radius in the limit of small chemical reaction amplitude A = ν−τ /∆c while keeping
the ratios ν−/(k−∆c) and k+/k− of reaction parameters fixed. This corresponds to the curves ¯R() shown in Fig. 2A
for different values of A. We can identify two regimes in the figure. The first is the region of small , ∼ 0, which
corresponds to the minimum of ( ¯R). The second is the region of ∞ where the stationary radius diverges. For A → 0,
we see that 0 goes to zero while ∞ stays constant, and both are connected by a straight line that indicates scaling
behavior of ¯R = ¯Rs. This increasing separation between 0 and ∞ (and the corresponding stationary radii) in the
limit of small A means that we can analyze the behavior of the stationary radius in these two regimes separately. For
this we consider Equations (A.2) and (A.3) for the concentration field and (A.6) for the stationary radius. We can
rewrite (A.6) to obtain an expression relating the supersaturation to the stationary radius,
(cid:18) β−γ
∆c ¯R
(cid:19) D−
ν−
+
k−∆c
D+
=
β+γ
∆c ¯R
+
l− − 1
¯R
l− coth ¯R
1 + ¯R
l+
.
(A47)
In this limit of small A, the characteristic length-scales of the concentration field become large with l± ∝ A−1/2. To
find scaling regimes in equation ((A47)), we change variables in Eq. (A47) from (A, ¯R) to (A, R) with R = ¯RAa/w,
where a is an exponent. For a = 1/3 we find the behavior of (R) close to 0 and ¯R0,
=
R−1 +
1
6
1
3
R2 + O(A1/6)
(A48)
where = A−1/3 becomes independent of A for small A. This function describes the supersaturation as a function
of radius around the threshold value 0. Due to the inverted presentation ( ¯R) instead of ¯R() the function captures
both the nucleation radius ¯Rc and the larger radius ¯Rs. The threshold value 0 can be obtained from Eq. (A48) by
minimizing for fixed A as ∂/∂ R = 0. It behave as
10
(A49)
For large and small R, Eq (A48) describes the steady radii ¯Rs and ¯Rc, respectively, for which ≥ 0. For large , the
critical radius obeys
.
0 = 4−2/3A1/3 + O(A1/2)
while the larger stationary radius is
¯Rc (cid:39) w
6
,
¯Rs (cid:39) w(3A)1/2 .
(A50)
(A51)
In Fig. 2B, the scaling behaviors given by Eq. (A51) and Eq. (A50) are indicated by dashed lines. At = 0 both radii
meet at ¯R = ¯R0, where
¯R0 = w(4A)−1/3 + O(A−1/2) .
For a = 1/2, ¯R/l± becomes independent of A and
=
D−
D+
ν−
k−∆c
(cid:16) ¯R
(cid:17) − 1
¯R
l− coth
l−
1 + ¯R
l+
+ O(A1/2)
(A52)
(A53)
For ¯R/l± (cid:28) 1, the stationary radius obeys Eq. (A51) and is thus the larger stationary radius ¯Rs. For ¯R/l± (cid:29) 1, we
obtain the divergence of ¯Rs as approaches ∞ with
(cid:115)
∞ =
D−k−
D+k+
ν−
k−∆c
.
(A54)
b. Shape instability
We now discuss scaling relations for the onset of instability in the (A, ) plane in the limit of small A, which give
the trends shown as dashed lines in Fig. 2D-F. We use the scaling of the stationary radius ¯R = ¯Rs close to 0 with
R = ¯RA1/3/w, = A−1/3 and l± = l±A1/2 in Eq. (13) to obtain
µnlm = − dl
R
A−2/3
F
+
2
3
(l − 1) − D+
D−
(l − 1)gl
R3
+ O(A1/6)
(A55)
where µnlm = µnlmτ /A and R is related to by (A48). Here, dl = fC3 − fC1, where fC1 and fC3 are defined in
Eq. (A33) and
gl =
with hl = (l2 + l − 2)/2. For large mode index l,
hl(l + 1) + D−
D+
l − 1
β−
β+
hll
dl =
l
2(η+/η− + 1)
+ O(1/l) .
(A56)
(A57)
We now consider conditions for which µnlm = 0 for small A and the mode (n, l, m) becomes unstable. Using (A48) in
(A55), we find a relation between and R at the onset of instability µnlm = 0,
(cid:18) 1
6
(cid:19)
=
dl
2(l − 1)
1
F
R +
+
D+
D−
1
2
gl
R−1 + O(A1/6) .
(A58)
11
This curve captures the scaling behavior of the onset of instability for different parameters in the ¯R − plane,
corresponding to the red dotted line in Fig. 2A-C.
We now focus on finding the scaling relations for the onset of stability of the stationary radius as function of A,
and F , as shown in Fig. 2D-F. At this onset, both (A58) and (A48) need to be satisfied. We use both equations to
eliminate R. We find a crossover regime with relations A∗ ∼ F −3/2 between the region where hydrodynamic flows are
relevant (A < A∗) and where they can be neglected (A > A∗). For A > A∗ we find for µnlm = 0 as relation between
A and
(cid:1)3 3 .
A (cid:39) 54
gl
2 gl
(cid:0)1 + 1
(cid:19)2
(cid:18) 2(l − 1)
(A59)
(A60)
For A < A∗ we find
A (cid:39) 1
3
−1F −2 .
dl
In Fig. 2D-F, the dashed lines indicate these two scaling solutions in the limit A → 0 and F → ∞ for l = 2, which
we find to be the first mode to become unstable. We find that the general trends of the stability diagram is captured
well, with small deviations from the full solution of Eq. (13) for small , and larger deviations in the regime close to
∞ where the scaling of the stationary radius ¯Rs ∝ A−1/3 breaks down.
Appendix B: Continuum model for active droplets with flows
1. Continuum model for active droplets
12
We study an extended Cahn-Hilliard equation with chemical reactions coupled to Stokes equation for hydrodynamic
flows at low Reynolds numbers. We consider an incompressible fluid containing two components A and B, with number
concentration fields cA(r, t) and c = cB(r, t) that depend on position r and time t, and with molecular masses mA
and mB and molecular volumes vA and vB. We are interested in the case where component A forms the background
fluid and B is a droplet material that forms droplets by phase separation. Additionally, chemical reactions convert the
two components into each other, A (cid:10) B. For simplicity, we consider mass and volume conserving chemical reactions
with mA/vA = mB/vB, which encodes that volume is conserved in the reaction if mass is conserved. Together with
incompressibility, this implies that the mass density ρ = mAcA + mBcB is constant. Therefore, we can describe the
system by the concentration c(r, t) of the droplet material B only.
We use the following double-well free energy density46
c − c(0)−
f (c) =
b
(cid:16)
(cid:17)2(cid:16)
(cid:17)2
c − c(0)
(cid:0)∇c(cid:1)2
κ
2
with ∆c =(cid:12)(cid:12)c(0)− − c(0)
+
(cid:12)(cid:12). Here, the positive parameter b characterizes molecular interactions and entropic contributions.
2(∆c)2
(B1)
+
+
,
This free energy describes the segregation of the fluid in two coexisting phases49: one phase rich in droplet material
with c ≈ c(0)− and a dilute phase with c ≈ c(0)
+ . The coefficient κ is related to surface tension and the interface width46.
The state of the system is characterized by the free energy
(cid:90)
F [c] =
d3r f (c) ,
(B2)
where the integral is over the system volume. We work with an ensemble T , ρ, c here, where T denotes temperature
and the system is considered isothermal. The chemical potential ¯µ = δF [c]/δc, governs demixing and can be split
into local and nonlocal contributions, ¯µ = ¯µ0 − κ∇2c with
(cid:0)c − c(0)
+
(cid:1)(cid:0)c − c(0)− (cid:1)(cid:0)2c − c(0)− − c(0)
+
(cid:1) .
¯µ0 =
b
(∆c)2
(B3)
(B4)
(B5)
The dynamics of the concentration field is described by50,51
∂tc = −∇ · j + s(c)
j = −M∇¯µ + vc .
Here, M is a mobility coefficient of the droplet material and v is the hydrodynamic velocity. The source term s(c)
describes chemical reactions, for which we choose for simplicity a linear concentration dependence,
s(c) = ν − k(c − c(0)
+ ) .
(B6)
The reaction flux given in Eq. (B6) does not obey detailed balance with respect to the free energy, and thus describes
a situation where an external energy source maintains the system away from equilibrium39.
The hydrodynamic velocity v can be calculated using momentum conservation,
with momentum ρvα and stress tensor σαβ, where α and β number cartesian coordinates x, y, z. We can write the
stress tensor σαβ as
∂t(ρvα) = ∂βσαβ ,
(B7)
σαβ = −(ρvα)vβ + σeq
αβ + σd
αβ ,
(B8)
where the first term describes advection of the stress tensor, σeq
tensors. The equilibrium stress tensor is given by
αβ and σd
αβ denote the equilibrium and dissipative stress
αβ = −(¯µc − f )δαβ − ∂f
σeq
∂(∂αc)
∂βc − P0δαβ .
(B9)
13
Here, ¯µc − f is the osmotic pressure of the droplet material, and δαβ denotes the Kronecker delta. Incompressibility
is enforced by an additional partial pressure P0. The deviatory stress tensor can be found as thermodynamic force
related to momentum by writing the entropy production rate,
(cid:18)
(cid:19)
σd
αβ = 2η
vαβ − 1
3
vγγδαβ
+ η(cid:48)vγγδαβ ,
(B10)
where η and η(cid:48) denote viscosities, and vαβ = (∂αvβ + ∂βvα)/2 is the symmetric strain tensor.
In the Stokes limit, the inertial terms are neglected, Dt(ρvα) = 0, with advected derivative Dt = ∂t + vβ∂β, leaving
0 = ∂β(σeq
αβ + σd
αβ). This yields52
βvα = 3¯µ0∂αc − κc∇2(∂αc) + ∂αP0 .
η∂2
(B11)
Eqs. (B3) -- (B6) and Eqs. (B11) and incompressibility ∂αvα = 0 define the continuum model of active droplets.
2. Numerical solution of the continuum model
We numerically solve the dynamic equations of the continuum model of active droplets, Eqs. (B3) -- (B6) and
Eqs. (B11) with Eq. (B13) and incompressibility ∂αvα = 0.
For this we use a spectral method in a 3d rectangular box. This has the advantage that in a spectral decomposition,
the spatial operators become simple multiplications with the wavenumber47. However, our equations contain a number
of nonlinear functions, which are easier to evaluate in real space. We therefore transform forward and back in each
time step.
To calculate the next timestep ti from the fields found in timestep ti−1, we use a semi-implicit Runge-Kutta
method53 (method (2,3,3)) for the concentration field. This evaluates the gradient term in ¯µ, Eq. (B3), implicitly,
while evaluating the rest of ¯µ as well as the advection term of the fluxes, vc, explicitly. This effectively means that
the terms related to the interfacial profile are calculated implicitly, which allows for larger time steps as an explicit
scheme.
For the concentration field, we choose no-flux boundary conditions (∂nc = 0, where the derivative is in a direction
normal to the simulation box), which leads to a decomposition in cosine functions in the spectral description. The
Laplacian then is −k2 for a mode with wave vector k. The Stokes equation can also be solved using spectral methods.
Here, no-flux conditions lead to vn = 0. Additionally we enforce incompressibility using a reprojection method. For
this, the velocity field calculated by neglecting the partial pressure, Pp = 0, can be split into two parts (Helmholtz
decomposition),
v = vψ + vφ = ∇ × ψ − ∇φ
with vector field ψ and scalar field φ, and velocity parts vψ = ∇ × ψ and vφ = −∇φ. With this, we find
(B13)
and thus, using incompressibility, ∇ · v = 0, we can calculate φ. We thus find the incompressible part of the velocity
field
∇ · v = ∆φ
vψ = v − ∇φ .
(B12)
(B14)
We can evaluate this in Fourier space using a spectral method. For a rectangular box aligned with the coordinate
system, we thus find that each velocity component vα is decomposed by sines in one direction and cosines in the other
direction. Spatial derivatives convert a sine-description into cosines, and vice versa.
We normalize concentration, length, time and energy by ∆c = c(0)− − c(0)
+ , w = 2(κ/b)1/2, t0 = w2/D and e0 =
κ w(∆c)2/3, respectively, where the characteristic length scale is w = 2(κ/b)1/2. The relevant dimensionless model
+ /∆c = 0, kt0 = 10−2, νt0 = 2 · 10−3 and η w3/(t0e0) = 2.
parameters are c(0)
Additionally, we use as box-length L/ w = 100 in all 3 dimensions, number of grid-points in one direction N = 128
and simulation time T /t0 = 4 · 103. For the time step, we start with a timestep of ∆t/t0 = 10−4, and double the
timestep to a final step size of ∆t/t0 = 0.01.
+ /∆c, kt0, and ν−t0/∆c. We choose c(0)
We start with initial conditions R = R0(1 + Y2,0), with R0/ w = 7 and = 1. The concentration field at positions
r is initialized by the function
c(r) =
c(0)
+ + c(0)−
2
+ − c(0)−
c(0)
2
+
tanh
d(r)
w
.
(B15)
where d(r) is the oriented distance of r to the nearest point on the ellipsoid. The value of d(r) is negative for points
inside the droplet and positive for points outside.
3. Effective droplet model as a limit of the continuum model
We now discuss the relationship between the effective droplet model and the continuum model. To relate the two
models, we first use the continuum model to derive jump conditions for the concentration in the effective droplet model
in equilibrium. We then consider stress balance across this interface and derive stress boundary conditions in the
effective droplet model. Finally we discuss the dynamical equations in the bulk and at the interface in non-equilibrium
situations.
14
a. Derivation of jump conditions for equilibrium phase separation
First we consider the phase separation in equilibrium without chemical reactions in the continuum model.
In a one-dimensional system with a mean concentration ¯c with c(0)
+ < ¯c < c(0)− , the free energy of the system in
Eq. (B2) is minimized by the concentration profile
c∗(x) =
c(0)− + c(0)
+
2
c(0)− − c(0)
+
2
+
tanh
x
w
,
(B16)
where w = 2(κ/b)1/2 denotes the interfacial width and x is the normal distance to the interface. The concentration
profile describes two phases of concentration c(0)− and c(0)
+ separated by a flat interface of width w. The surface tension
can be defined as
For the free energy Eq. (B2) with the concentration profile Eq. (B16), this can be written as γ = (cid:82) ∞
(F [c(0)− ] + F [c(0)
+ ])dx .
γ =
F [c∗(x)] − 1
2
(B17)
−∞ κ(∇c∗)2dx
which yields γ = (∆c)2/6
κb or54.
√
(cid:90) ∞
−∞
This interfacial tension governs the concentration jump condition in the effective droplet model, which can be
derived as follows. To describe a curved interface, we consider two homogeneous phases with concentrations c±. For
a finite volume Vs with a droplet of size V and area A the concentrations c± can be found by minimizing the free
energy F = f (c−)V + f (c+)(Vs − V ) + γA with ∂F/∂c−V = 0 and ∂F/∂V c− = 0, where the concentration of both
phases are related by Vs¯c = V c− + (Vs − V )c+ where ¯c denotes the average concentration in the system. Thus for
two phases to be in equilibrium, their chemical potential ¯µ and osmotic pressure Π = c¯µ − f need to obey
0 = ¯µ(c−) − ¯µ(c+)
0 = Π(c+) − Π(c−) − 2γH ,
(B18)
(B19)
where H the mean curvature of the droplet and 2γH is the Laplace pressure. These equations determine the concen-
trations in the phases c± of coexisting phases54.
For small Laplace pressures, we can express the equilibrium concentrations c± of a curved interface by the concen-
trations of a flat interface c(0)± plus a small perturbation,
(B21)
where β± = 2/(f(cid:48)(cid:48)(c(0)± )∆c). For the free energy Eq. (B2), we find β± = 2/(b∆c), which is related to the interfacial
width as w = 6γβ+/∆c.
c− = c(0)− + β−γH
c+ = c(0)
+ + β+γH
(B20)
b. Stress balance across the interface
We now consider stress balance of the continuum model across the droplet interface to derive stress jump conditions
at the interface in the effective droplet model. We discuss the mechanical equilibrium in a small volume across a curved
interface with a local mean curvature H corresponding to a (local) effective radius R = 1/H. We focus on the case
where the interface is rotationally symmetric around the considered point R, and where the curvature does not change
along the interface. We use spherical coordinates, where the radial vector er is aligned with the (outward pointing )
normal vector n and the tangential vectors t and s are aligned with eθ and eφ, respectively (with the vector directions
15
FIG. 4. Geometry for the force balance. We consider a spherical cap of the droplet interface, with a box with constant
distance δ to the interface inside and outside. The normal and tangential vectors n, t and s of the interface are shown, as well
as the normal vector n of the box. The origin of the spherical coordinate system is the center of the sphere that describes the
interfacial curvature, with radius R, while θ0 gives the polar angle of the cap.
for φ = 0 in the limit θ = 0). We consider a small box enclosing R where the outer and inner surfaces Aout and Ain
have a constant distance of δ to the interface, and the lateral surface Alat is at a constant angle θ0 with respect to
the symmetry axis. The geometry is shown in Fig. 4.
Now let us consider the balance of the stress tensor Eq. (B7) across the box, taking into account the curved geometry.
The stress balance ∂βσαβ can be written as
0 =
dA nβσαβ
(B22)
(cid:73)
(cid:90)
(cid:90)
(cid:90)
where α and β are cartesian coordinates and n the (local, outward pointing) normal vector of the box-surface. We
can split this in three terms,
0 =
dAoutσαn −
dAinσαn +
dAlatσαt ,
(B23)
where we used that the orientation of the normal vectors of the box coincides with the normal/tangential vector of
the interface.
On the inner and outer areas Ain and Aout, the stress tensor presented in Eq. (B8) with equilibrium stress tensor
in Eq. (B9) reduces to the form of the effective droplet model given after Eq. (6) in the main text, as the gradient
terms are negligible for δ (cid:29) w. We now consider the limit of a sharp interface w → 0 with finite surface tension γ,
and consider the case of a small box of thickness δ, which remains larger than the interfacial width. The components
α = x, y of Eq. (B23) vanish by symmetry. For α = z we find
where σ±
Integration over the lateral box surface Alat yields the last term,(cid:82) dAlatσαt
nn are the stress tensor components of the effective model, Eq. (4), inside and outside the interface at R.
∼= 2π R sin2 θ0 γ. We thus find that the
0 = π R2 sin2 θ0 σ+
nn − π R2 sin2 θ0 σ−
nn − 2π R sin2 θ0 γ ,
(B24)
mechanical equilibrium of a curved interface introduces a Laplace pressure 2γH,
0 = σ+
nn − σ−
nn − 2γH .
(B25)
We therefore recover the stress jump conditions of the effective droplet model, Eq. (6). Additionally, (B25) together
with (B19) implies that the partial pressure needed to satisfy incompressibility is continuous across the interface,
0 = P −
P +
0 .
c. Dynamics of the effective droplet model
16
We now consider the dynamics of a non-equilibrium system with a droplet. We show how the continuum model is
related to the bulk equations and jump conditions of the effective droplet model. For this we consider a droplet with
a interface that is thin compared to the dynamical length scales l±, so that we can describe the interface by local
equilibrium. In the bulk phases we focus on the case where deviations from the equilibrium concentrations are small.
In the bulk phases, we expand the chemical potential Eq. (B3) around the reference concentrations c(0)± . The
gradient term −κ∇2c in the chemical potential is important within the interface, but can be ignored in the bulk
phases, where the length-scales on which the concentration field varies are much larger than the interfacial width.
Thus we can describe the chemical potential by
¯µ±(c) ≈ d¯µ0
dc
(c − c(0)± ) ,
(B26)
(cid:12)(cid:12)(cid:12)(cid:12)c(0)±
which is ¯µ±(c) ≈ b(c − c(0)± ) for our specific free energy. With this simplification, Eqns.
(B4) and (B5) become
the reaction-diffusion-convection equations (1) and (2) with diffusion constants D± = M (d¯µ0/dc)c(0)±
or D± = M b.
Similarly we linearize the chemical reaction rate Eq. (B6) in both phases. As we already chose a linear rate for the
continuum model, we only need to relate the parameters k and ν with the constants k± and ν± of the effective model,
with k± = k, ν+ = ν and ν− = k∆c − ν. Inserting the linearized chemical potential Eq. (B26) into the equilibrium
stress tensor (B9) we find that momentum conservation in the bulk phases is given by the Stokes equation (4) with
viscosities η± = η, where the pressure p is determined by the incompressibility condition ∂αvα = 0.
We consider the droplet interface to be in local equilibrium. We therefore obtain Eq. (8) for the jump of the
concentration field in the effective model. The incompressibility condition ∂αvα = 0 implies v−
n (R) at a
sharp interface, and we consider an interface without slip length, so that v−(R) = v+(R). We thus find Eq. (7) of the
effective model. The normal stress balance in Eq. (6) is derived in B 3 b.
n (R) = v+
As a last point we need to find Eq. (10) for the interface movement. We consider the concentration change in
a box of width δ around the interface, see Fig. 4. We consider a box enclosing a point R on the interface at the
time t aligned with the normal and tangential directions of the interface at R. The interface may move with normal
movement ∂t R(t), with R(t) = R(t)· n and normal vector n, while the box stays at a fixed position. The total change
of material in the volume is given by
(cid:90)
(cid:90)
dV c = −
dA n · j +
∂t
V
A
V
(cid:90)
dV s(c)
(B27)
where V denotes the volume and A the area of the box. For small w and finite δ the concentration field c makes a
jump from the surface Ain to Aout given by conditions (8) and (9) at R. Within each phase, we can express the field
by the boundary values at the interface Eq. (B21) and a linear expansion,
c(r, t) (cid:39)
c−(R(t)) + ∇c−(r, t) · (r − R(t))
c+(R(t)) + ∇c+(r, t) · (r − R(t))
inside droplet
outside droplet
(B28)
The chemical reaction is given in both phases by Eq. (B6). For small δ and θ0, we find for the left-hand side of
Eq. (B27) that δc vanishes to lowest order and
dV c = AR(c−(R(t)) − c+(R(t)))∂t R + O() + O(θ0)
(B29)
for a small box,(cid:82)
where AR is the area of the droplet interface enclosed by the box. For a spherical cap, AR = 2π(1 − cos θ0) R2. We
further find that the source term due to the chemical reaction scales with the volume of the box, and thus vanishes
V dV s(c) = 0 + O() + O(θ0). The flux across the box can be expressed as
−
dA n · j = ARn · (j−(R(t)) − j+(R(t))) + O() + O(θ0)
(B30)
where j±(R(t)) denotes the flux at R inside/outside the droplet. We thus find the normal movement of the interface,
∂t R = n · j−(R(t)) − j+(R(t))
c−(R(t)) − c+(R(t))
.
(B31)
(cid:40)
(cid:90)
∂t
V
(cid:90)
A
17
In the main text we use spherical coordinates centered at the droplet center. For a spherical droplet, the normal
and radial movement would thus be the same. For a deformed droplet, we need to consider the relation between the
normal interface movement, R(t) = R(t) · n and the radial movement R(t) = R(t) · er. At fixed angles θ and φ,
the interface movement is given by ∂tR = ∂tR er. Using ∂t R = ∂tR(t) · n, we find a relation between the radial
and normal movement, ∂tR = ∂t R/(n · er). This relation, together with Eq. (B31), yields the interfacial movement
Eq. (10) presented in the main text.
We thus recover all dynamical equations of the effective droplet model from the continuum model based on irre-
versible thermodynamics. Note that the specific choice of the free energy leads to specific relations between parameters
of the effective model such as D+ = D−. Our derivation shows the relation between both models in the case where
the interface width w is small compared to the droplet size, R/w (cid:29) 1, and the chemical diffusion length, l±/w (cid:29) 1.
Additionally, we focused on the case where the concentrations in the phases are similar to the concentrations in equi-
librium and have small concentration gradients. These conditions are not valid in all systems. Most importantly, the
chemical reactions can drive concentrations far away from the equilibrium phase concentrations c(0)± . The resulting
behaviors, such as the formation of new interfaces associated with instabilities of the spinodal decomposition regime,
are not captured in the effective droplet model.
18
FIG. 5. Growth of shape perturbations of the l = 2 mode for different normalized viscosities F = ηw/(γτ ) for the continuous
model (red crosses) and effective model (blue curve). The last data point (with arrow) corresponds to F → ∞. (Parameters:
A = 8 · 10−3, = 0.2, η−/η+ = 1, c(0)
+ /∆c = 0, k+/k− = 1, ν−/(k−∆c) = 0.8)
4. Comparison of the droplet dynamics in the continuum model and the effective model
Here we compare the analytical predictions of the effective model for the instability with numerical calculations of
the continuous model for different values of the renormalized viscosity F . For this we numerically solved the dynamic
equations of the continuous model starting with a droplet with a small initial deformation of mode l = 2. We fitted
the dynamical behavior of the mode to an exponential function, with yields a numerical estimate for the eigenvalue
µ2. In figure 5 the resulting eigenvalues are shown, together with the eigenvalue of corresponding parameters of the
effective model. We find that the value of F for which droplet shapes become unstable is very similar to the value
predicted by the effective model. The eigenvalues are qualitatively similar to the ones of the effective model, despite
working in an a parameter regime where the interfacial width and the differences of concentration within a phase
cannot be considered very small, so that the models are not necessarily comparable.
To generate the data in the figure, we initialized droplets with a small shape perturbation for different values of
F . All parameters and initial conditions were chosen as described in B 2. We found that for F ≥ 100 droplets divide,
while they are stable for F ≤ 1. For F = 10, the shape deformation was very slow, so that division was not seen in the
time interval T /τ = 4000. For 10 < F < 100, as well as F = ∞, we fitted radius and spherical harmonic deformation
to the concentration field using Eq. (B15). For short times, the droplet radius changes as the concentration field and
droplet size go towards the stationary values. After that, the shape deformation grows until the droplet deforms
so strongly that the fitting fails. By hand we chose intermediate time windows for the simulations where the size
was stationary and the shape deformation small. In these windows we fitted the deformation amplitude (compare
Eq. (B15)) with an exponential function, Aeµ2t + B with parameters A, B and eigenvalue µ2 to the l = 2 mode of the
shape deformation.
Dimensionless viscosityDimensionless eigenvalueAppendix C: Estimation of parameters
19
Here we estimate the hydrodynamic parameter for two physical phase-separating systems to understand the impor-
tance of hydrodynamic flows on the droplet division in experimental systems. We discuss two cases, water-oil phase
separation, and soft colloidal systems (such as protein-RNA phase-separation in cells). We have already estimated pa-
rameter values for both systems without the influence of hydrodynamic flows39, where we found that droplet division
should be possible for realistic values of chemical reaction rates in both systems, and that corresponding stationary
radii would have sizes of a few micrometers. Here we estimate the value of the dimensionless viscosity F for water-oil
and soft colloidal systems, and compare them to the analytical phase diagrams presented in Fig. 2.
To calculate the hydrodynamic parameter F for experimental systems, we need an estimation of the diffusion
coefficient of the droplet material D+ outside the droplet, of the interfacial width w (which corresponds to length-
scale w in the paper39), of the surface tension γ and of the viscosity η− inside the droplet. For water-oil systems,
the interfacial width is of the order of w ≈ 1nm and the diffusion constant is D+ ≈ 10−9m2/s. We can estimate the
surface tension as γ ≈ 10−2N/m, and the viscosity η− ≈ 10−3(N · s)/m254,55. With these values, we find F ≈ 0.1. In
this case droplet division is strongly suppressed, see Fig. 2 of the main text. For soft colloidal systems, we estimate
w ≈ 10nm, D+ ≈ 10−10m2/s and γ ≈ 10−6N/m1,54. The value of F depends on the viscosity of the droplet. For
values η− ≈ 10−3(N · s)/m2, F ≈ 10, and for η− ≈ 1− 10(N · s)/m2, we have F ≈ 104. In both cases droplet division
is possible, but more easy to achieve for larger F . We convert A∗ to the reaction rate ν− inside the droplet using the
droplet concentration given in39.
We can use Eq. (A59) and (A60) from the scaling analysis to estimate the instability of the concrete parameter
examples discussed in39 under the influence of hydrodynamic flows. In these scaling equations, the ratios η+/η− and
D−β−/(D+β+) enter the calculation of A∗ and ∗ but we find that they do not lead to relevant changes in the results.
The scaling analysis thus yields results very similar to the estimation using Fig. 2.
20
1 C. P. Brangwynne, C. R. Eckmann, D. S. Courson, A. Rybarska, C. Hoege, J. Gharakhani, F. Julicher, and A. A. Hyman,
Science 324, 1729 (2009).
2 C. P. Brangwynne, Soft Matter 7, 3052 (2011).
3 A. A. Hyman and K. Simons, Science 337, 1047 (2012).
4 P. Li, S. Banjade, H.-C. Cheng, S. Kim, B. Chen, L. Guo, M. Llaguno, J. V. Hollingsworth, D. S. King, S. F. Banani, P. S.
Russo, Q.-X. Jiang, B. T. Nixon, and M. K. Rosen, Nature 483, 336 (2012).
5 S. C. C. Weber and C. P. P. Brangwynne, Cell 149, 1188 (2012).
6 C. P. Brangwynne, J. Cell Biol. 203, 875 (2013).
7 J. a. Toretsky and P. E. Wright, The Journal of Cell Biology 206, 579 (2014).
8 D. Zwicker, M. Decker, S. Jaensch, A. A. Hyman, and F. Julicher, Proc. Natl. Acad. Sci. USA 111, E2636 (2014).
9 S. Elbaum-Garfinkle, Y. Kim, K. Szczepaniak, C. C.-H. Chen, C. R. Eckmann, S. Myong, and C. P. Brangwynne, Proceedings
of the National Academy of Sciences 112, 201504822 (2015).
10 A. Molliex, J. Temirov, J. Lee, M. Coughlin, A. P. Kanagaraj, H. J. Kim, T. Mittag, and J. P. Taylor, Cell 163, 123 (2015).
11 A. Patel, H. O. Lee, L. Jawerth, S. Maharana, M. Jahnel, M. Y. Hein, S. Stoynov, J. Mahamid, S. Saha, T. M. Franzmann,
A. Pozniakovski, I. Poser, N. Maghelli, L. A. Royer, M. Weigert, E. W. Myers, S. Grill, D. Drechsel, A. A. Hyman, and
S. Alberti, Cell 162, 1066 (2015).
12 M. Feric, N. Vaidya, T. S. Harmon, D. M. Mitrea, L. Zhu, T. M. Richardson, R. W. Kriwacki, R. V. Pappu, and C. P.
Brangwynne, Cell 165, 1686 (2016).
13 S. Saha, C. A. Weber, M. Nousch, O. Adame-Arana, C. Hoege, M. Y. Hein, E. Osborne-Nishimura, J. Mahamid, M. Jahnel,
L. Jawerth, A. Pozniakovski, C. R. Eckmann, F. Ju, and A. A. Hyman, Cell 166, 1 (2016).
14 S. F. Banani, H. O. Lee, A. A. Hyman, and M. K. Rosen, Nature Reviews Molecular Cell Biology 18, 285 (2017).
15 E. Sokolova, E. Spruijt, M. M. K. Hansen, E. Dubuc, J. Groen, V. Chokkalingam, A. Piruska, H. A. Heus, and W. T. S.
Huck, Proceedings of the National Academy of Sciences 110, 11692 (2013).
16 Y. Lin, D. S. W. Protter, M. K. Rosen, and R. Parker, Molecular Cell 60, 1 (2015).
17 W. M. Aumiller and C. D. Keating, Nature Chemistry 8, 129 (2015).
18 E. A. Frankel, P. C. Bevilacqua, and C. D. Keating, Langmuir 32, 2041 (2016).
19 K. K. Nakashima, J. F. Baaij, and E. Spruijt, Soft Matter 14, 361 (2018).
20 S. C. Glotzer, D. Stauffer, and N. Jan, Phys. Rev. Lett. 72, 4109 (1994).
21 S. Puri and H. Frisch, J. Phys. A 27, 6027 (1994).
22 J. J. Christensen, K. Elder, and H. C. Fogedby, Phys. Rev. E 54, R2212 (1996).
23 D. Carati and R. Lefever, Phys. Rev. E 56, 3127 (1997).
24 A. Z. Patashinski, R. Orlik, K. Paclawski, M. A. Ratner, and B. A. Grzybowski, Soft Matter 8, 1601 (2012).
25 L. Giomi and A. DeSimone, Phys. Rev. Lett. 112, 147802 (2014).
26 D. Zwicker, A. A. Hyman, and F. Julicher, Phys. Rev. E 92, 012317 (2015).
27 J. Crosby, T. Treadwell, M. Hammerton, K. Vasilakis, M. P. Crump, D. S. Williams, and S. Mann, Chemical Communications
48, 11832 (2012).
28 T.-Y. Dora. Tang, D. van Swaay, A. DeMello, J. L. Ross Anderson, and S. Mann, Chem. Commun. 51, 11429 (2015).
29 A. I. Oparin, Izd.Moskovhii RabochiI (1924).
30 A. I. Oparin, Origin of Life (Oliver and Boyd, Edinburgh, 1957).
31 S. W. Fox, Orig. Life 7, 49 (1976).
32 M. M. Hanczyc and J. W. Szostak, Curr. Opin. Chem. Biol. 8, 660 (2004).
33 K. P. Browne, D. A. Walker, K. J. M. Bishop, and B. A. Grzybowski, Angew. Chem. 49, 6756 (2010).
34 S. Koga, D. S. Williams, A. W. Perriman, and S. Mann, Nature Chemistry 3, 720 (2011).
35 G. Murtas, Mol. BioSyst. 9, 195 (2013).
36 A. A. Hyman, C. A. Weber, and F. Julicher, Annual Review of Cell and Developmental Biology 30, 39 (2014).
37 M. Li, X. Huang, T. Y. D. Tang, and S. Mann, Current Opinion in Chemical Biology 22, 1 (2014).
38 T.-Y. Dora Tang, C. Rohaida Che Hak, A. J. Thompson, M. K. Kuimova, D. S. Williams, A. W. Perriman, and S. Mann,
Nature Chemistry 6, 527 (2014).
39 D. Zwicker, R. Seyboldt, C. A. Weber, A. A. Hyman, and F. Julicher, Nature Physics 13, 408 (2017).
40 S. Lach, S. M. Yoon, and B. A. Grzybowski, Chem. Soc. Rev. 65, 1392 (2016).
41 L. Rayleigh, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 34, 145 (1892).
42 S. Chandrasekhar, Hydrodynamic and hydromagnetic stability (Dover Publications, Inc., New York, 1981).
43 P. Constantin, T. F. Dupont, R. E. Goldstein, L. P. Kadanoff, M. J. Shelley, and S. M. Zhou, Physical Review E 47, 4169
(1993).
44 J. D. Paulsen, R. Carmigniani, A. Kannan, J. C. Burton, and S. R. Nagel, Nature Communications 5, 3182 (2014).
45 R. G. Barrera, G. a. Estevez, and J. Giraldo, European Journal of Physics 6, 287 (1985).
46 J. W. Cahn and J. E. Hilliard, The Journal of Chemical Physics 28, 258 (1958).
47 L. Q. Chen and J. Shen, Computer Physics Communications 108, 147 (1998).
48 J. Stenesh, Biochemistry (Springer US, 2013).
49 R. C. Desai and R. Kapral, Dynamics of Self-organized and Self-assembled Structures (Cambridge University Press, 2009).
50 S. R. De Groot and P. Mazur, Non-equilibrium thermodynamics (Dover Publications, Inc., New York, 2011).
51 J. Prost, F. Julicher, and J.-F. Joanny, Nature Physics 11, 131 (2015).
52 M. E. Cates, Soft Interfaces: Lecture Notes of the Les Houches Summer School: Volume 98, July 2012 98, 317 (2017).
53 U. Ascher, S. Ruuth, and R. Spiteri, Applied Numerical Mathematics 25, 151 (1997).
54 S. A. Safran, Statistical thermodynamics of surfaces, interfaces, and membranes, Frontiers in Physics, Vol. 90 (Addison-
Wesley, Reading, 1994).
55 W. M. Haynes, CRC handbook of chemistry and physics (CRC press, 2014).
21
|
1801.02464 | 2 | 1801 | 2018-04-16T12:06:26 | Fluid flows shaping organism morphology | [
"physics.bio-ph",
"q-bio.TO"
] | A dynamic self-organized morphology is the hallmark of network-shaped organisms like slime moulds and fungi. Organisms continuously re-organize their flexible, undifferentiated body plans to forage for food. Among these organisms the slime mould Physarum polycephalum has emerged as a model to investigate how organism can self-organize their extensive networks and act as a coordinated whole. Cytoplasmic fluid flows flowing through the tubular networks have been identified as key driver of morphological dynamics. Inquiring how fluid flows can shape living matter from small to large scales opens up many new avenues for research. | physics.bio-ph | physics | Fluid flows shaping organism morphology
1Max Planck Institute for Dynamics and Self-Organization, Am Fassberg 17, D-37077 Gottingen, Germany∗
Karen Alim1
A dynamic self-organized morphology is the hallmark of network-shaped organisms like slime
moulds and fungi. Organisms continuously re-organize their flexible, undifferentiated body plans to
forage for food. Among these organisms the slime mould P. polycephalum has emerged as a model
to investigate how organism can self-organize their extensive networks and act as a coordinated
whole. Cytoplasmic fluid flows flowing through the tubular networks have been identified as key
driver of morphological dynamics. Inquiring how fluid flows can shape living matter from small to
large scales opens up many new avenues for research.
Many organisms, including a broad range of species of
slime moulds and fungi, grow and forage as a single large
network. Networks change their morphology over and
over again during growth and migration to locate food
in a patchy environment [1, 2], see also Fig. 1. More-
over, network morphology adapts in response to newly
acquired food sources, even connecting food sources in
an efficient and robust manner [3]. The striking similar-
ity of the morphological dynamics in foraging fungi and
slime moulds is even more surprising if one takes into
account that slime moulds and fungi are genetically dis-
tinct, with slime moulds being genetically even closer to
animals than fungi.
It is therefore likely that not bio-
logical make-up but the physics of fluid flows within the
tubular networks are critical to the self-organization of
network morphology across an individual. Both kinds of
living, adaptive networks exhibit oscillatory, long-ranged
fluid flows [4–6]. Here, the syncytial plasmodia of the
slime mould Physarum polycephalum emerged as a model
system to understand the role of flows in coordinating
morphology. Fluid flows in this organism are highly co-
ordinated, driving intracellular transport on short time
scales but also migration and likely morphological self-
organization at long time scales.
CYTOPLASMIC FLOWS ORGANIZED IN
PERISTALTIC WAVE
In P. polycephalum the fluid cytoplasm within the
tubular network streams forth and back in a shuttle flow
[6, 7]. Network sizes are macroscopic ranging from about
500 µm to 0.5 m - experiments are typically conducted on
specimen of up to few cm in size. Flows generally exhibit
a Poiseuille profile [6, 8] with deviations likely in smaller
tubes [9]. Flow is dominated by small Reynolds num-
ber Re = 2U R/ν ≈ 0.002 and small Womersley num-
ber α = R(cid:112)ω/ν ≈ 0.004, based on a representative
tube radius of R = 50 µm, a flow velocity reaching up
to U = 1 mm/s [6], a kinematic viscosity of cytoplasm
ν = 6.4 × 10−6 m2/s [10] and an oscillation frequency of
ω = 0.05 Hz. The cytoplasm is enclosed by gel-like walls
that are lined with an actin cortex [11, 12]. Actin orga-
nizes in circumferential fibrils that contract periodically
[13] and drive the cytoplasms' shuttle flow. Contrary
to long-lasting speculations about localized pumps driv-
ing pressure difference, the common understanding now
is that flows arise through network-wide, self-organized
contractions of the actin cortex [14, 22].
The shuttelling cytoplasm itself is very rich in actin.
In 1 mm sized cytoplasm extract droplets, so called
proto-plasmic droplets, the actin cortex and the con-
tractions self-organize over time, showing first irregular
contraction patterns and later highly coordinated spatio-
temporal patterns including standing, travelling and spi-
ral waves [15–18]. Similar dynamics and patterns can
be reproduced in models by describing the proto-plasmic
droplets by two phases of a viscoelastic solid phase rep-
resenting the cytoskeleton, interpenetrated by a second
fluid phase representing the cytosol and coupling both
phases by a soluble molecule that activates tension in
the solid phase [19, 20]. Even on the much larger scale
of an entire tubular network contractions are also highly
coordinated, forming a peristaltic wave spanning speci-
men of at least 2 cm in size [21, 22], see Fig. 1. While the
contraction period only increases moderately with net-
work size [24, 25], the wavelength of the peristaltic wave
matches organism size spanning two orders of magnitude
[22]. It is fascinating to investigate how this scaling can
arise. Given the success of mechanochemical models for
proto-plasmic droplets it is likely that also the organism
spanning peristaltic wave is a result of mechanochemical
patterning, which could nicely complement our emerg-
ing understanding of this novel patterning mechanism
widespread in biological systems [26, 27]. From a fluid
mechanics point a view a peristaltic wave matching or-
ganism size induces the highest flow velocities through-
out a network. As flows change with organism morphol-
ogy, they are likely not only important for organism ho-
moeostasis but also for the coordination of morphological
adaptation itself.
EFFECTIVE INTRACELLULAR TRANSPORT
BY OSCILLATORY FLOW
Peristalsis is a common mechanism in biological sys-
tems, creating oscillatory flow and pumping fluid along
8
1
0
2
r
p
A
6
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
4
6
4
2
0
.
1
0
8
1
:
v
i
X
r
a
2
FIG. 1.
Network morphology and peristaltic wave driving long-ranged cytoplasmic flows in the plasmodial slime mould
P. polycephalum (a) Self-organization of the network morphology over the course of 12.5 h. An initially well-reticulated network
migrates to the left thereby retracting connections and altering tube radius hierarchy. Scale bar, 5 mm.
(b) Bright-field
microscopy image of P. polycephalum network trimmed to a rectangular shape. Scale bar, 2 mm.
(c) Analysis of actin-
cytoskeleton-driven contractions of tube walls in (b) reveal phases of contractions to be patterned in a roughly linear gradient.
The wavelength of this peristaltic phase pattern on a network matches the organism size. (d) Simulation results of the spread
of particles (copper colour) due to the streaming of fluid flow driven by the peristaltic contractions.
a tube [28–31]. In P. polycephalum the tubular network
can be considered to be of fixed volume on the time scales
of tens of contraction periods. Therefore, net fluid trans-
port is not relevant on these short time scales [22, 32].
Yet, creating shuttle flows by a peristaltic wave of con-
tractions is a simple but powerful mechanism to increase
the spread of any particles, like metabolites or signalling
molecules within this closed network. Rare measure-
ments show organism-wide ∼ 2 cm transport of particles
within half a contraction period [21] out-competing dif-
fusive spread that would have travelled only 0.25 mm in
that time frame. Because of the oscillatory nature of
flow, particles flow mainly back to their initial site after
a whole period. But the peristaltic flow also increases
the effective diffusion κ → κ + U 2R2/48κ according to
Taylor dispersion in long slender tubes [33, 34] also ap-
plicable to contractile tubes [35, 36]. Here, κ denotes
the bare molecular diffusivity. Rapid diffusion across the
tiny tube cross-sections allows particles to transition be-
tween fast and slow streamlines of the Poiseuille profile
rapidly increasing their dispersion along the tube by an
order of magnitude, see Fig. 2 c. The adjacency of big
and small tubes and therefore different flow velocities in
a slime mould network make it a hard task to theoret-
ically map out how far particles can spread [37]. One
successful strategy is to map out an effective dispersion
[38, 39], which unveils for example that particle spread
is increased by pruning/coarsening of the network when
P. polycephalum is left to starve [39]. This already pro-
vides a glimpse of the challenging question on how net-
work morphology impacts network-wide transport [21].
Is morphology geared to optimizing dispersion? If so,
how does it evolve toward an optimized morphology?
Transport by fluid flow seems to lie at the basis of net-
work self-organization since advected signaling molecules
may propagate information about the acquisition of a
food source throughout the network by hijacking fluid
flows [40]. Here, signaling molecules advected by fluid
flow directly increase contraction activity. Further, flow
is necessary to synchronize and coordinate contractions
[41–43]. It is therefore likely that flows play a crucial role
in coordinating contractions over space and time.
0h6:15h12:30hadbcCELL MIGRATION BY PUMPING OF
PERISTALTIC WAVE
Cytoplasmic flows form the basis of fast locomotion
of very different kinds of amoeba [44–47]. Flows arise
by local expansion of the actin cortex and subsequent
myosin dependent contraction [45]. Flows and contrac-
tility underlying cell migration are well accessible in the
amoeboid plasmodia of P. polycephalum. Plasmodia of
100 µm to 500 µm length adopt an amoeboid shape mi-
grating rapidly [48, 49] by net fluid transport generated
by a contractile peristaltic wave and cortex expansion at
the front [50]. Changes in contraction pattern affect lo-
comotion velocity [51]. Plasmodia that exceed ∼ 500 µm
form a full network structure often including multiple
migration fronts. Over the peristaltic cycle a front lo-
cation advances and retracts asymmetrically leading to
net advancement at long time scales [52]. As networks
grow they move faster [24]. When confined to lanes, mi-
gration velocity scales linearly with maximal plasmodium
height, reaching locomotion speeds of up to v = 0.4mm/s
[24]. Reorientation of migration direction for example to-
ward a food source is associated with a redirection of the
peristaltic wave direction to that site [40] further sub-
stantiating that migration is governed by fluid flows on
long time scales. Given that signaling molecules are ad-
vected by fluid flows affecting actin cortex dynamics [40]
it is seems possible that flows play an important role in
the navigation of organisms, acting for example during
chemotaxis.
MORPHOLOGICAL CHANGES TRIGGERED BY
CYTOPLASMIC FLOWS
Fluid flows not only transport particles and fluid mass
but also exert forces themselves that may induce long-
term changes to morphology. Forces may directly feed
back onto biochemical reactions triggering complex spa-
tiotemporal dynamics due to this mechanochemical cou-
pling as currently more and more observed in morpho-
genetic processes [26, 27]. Even without the ability to
pin down a specific feedback on chemical reactions the
influence of forces generated by flow can be investigated
on a coarse-grained level. For example, in animal vas-
culature it is observed that tube diameters grow with
increased flow rate regulating shear force to a balanced
level [53]. This observation inspired the idea that fluid
shear force induces morphological changes in vasculature,
a concept also successfully used in models of P. poly-
cephalum dynamics [21]. Further support is the success
of Murray's law particularly in plant and animal vascu-
lature [54] which predicts the ratio of tube diameters at
a network node under the assumption of conserved shear
force. Murray's law is also consistent with minimizing
dissipation inspiring theoretical work on optimal network
3
architectures [55–57]. Given that P. polycephalum grows
its almost transparent tubes in a planar network, testing
principles such as Murray's law [58] and in general re-
lating morphological dynamics to flow properties is very
feasible.
In particular, the adaptability of the network
morphology makes it a very suitable system to explore
how well certain properties like dissipation, robustness
[3] or transport capabilities [59, 60] are optimized by liv-
ing organisms. Equally, as flows are globally coupled
throughout the network, there is considerable additional
complexity in this system. Indeed, it might well be that
precisely this added complexity due to the coupling is
the key to have simple mechanisms based on fluid flow
give rise to the complex dynamics of self-organization of
morphology we observe.
ACKNOWLEDGEMENT
This work has been supported by the Max Planck So-
ciety
∗ [email protected]
[1] L Boddy, J Hynes, D P Bebber, and M D Fricker. Sapro-
trophic cord systems: Dispersal mechanisms in space and
time. Mycoscience, 50:9–19, 2009.
[2] S L Stephenson and H Stempen. Myxomycetes: a hand-
book of slime molds. Timber Press, Inc., Oregon, 1994.
[3] A Tero, S Takagi, T Saigusa, K Ito, D P Bebber, M D
Fricker, K Yumiki, R Kobayashi, and T Nakagaki. Rules
for biologically inspired adaptive network design. Sci-
ence, 327:439–42, 2010.
[4] M Tlalka, S C Watkinson, P R Darrah, and M D Fricker.
Continuous imaging of amino-acid translocation in intact
mycelia of Phanerochaete velutina reveals rapid, pulsatile
fluxes. New Phytol., 153:173–184, 2002.
[5] M Tlalka, D P Bebber, P R Darrah, S C Watkinson,
and M D Fricker. Emergence of self-organised oscillatory
domains in fungal mycelia. Fungal Genet. Biol., 44:1085–
1095, 2007.
[6] N Kamiya. The rate of the protoplasmic flow in the myx-
omycete plasmodium I. Cytologia, 15:183–193, 1950.
[7] P A Stewart and B T Stewart. Protoplasmic movement
in slime mold plasmodia: The diffusion drag force hy-
pothesis. Exp. Cell Res., 17:44–58, 1959.
[8] Alexander V Bykov, Alexander V Priezzhev, Janne
Lauri, and Risto Myllyla. Doppler OCT imaging of cy-
toplasm shuttle flow in Physarum polycephalum. J. Bio-
photon., 2(8-9):540–547, 2009.
[9] Yurii Romanovskii and VA Teplov. The physical bases
of cell movement. The mechanisms of self-organisation of
amoeboid motility. Phys.-Usp., 38(5):521–542, 1995.
[10] R Swaminathan, C P Hoang, and A S Verkman. Pho-
tobleaching recovery and anisotropy decay of green flu-
orescent protein GFP-S65T in solution and cells: cy-
toplasmic viscosity probed by green fluorescent protein
translational and rotational diffusion. Biophys. J., 72(4):
1900–1907, 1997.
[11] G Isenberg and K sE Wohlfarth-Bottermann. Transfor-
mation of cytoplasmic actin. Importance for the organi-
zation of the contractile gel reticulum and the contraction
– Relaxation cycle of cytoplasmic actomyosin. Cell Tiss.
Res., 173:495–528, 1976.
[12] W Naib-Majani, W Stockem, K Weber, J Wehland, and
K E Wohlfarth-Bottermann. Cytoplasmic actin patterns
in Physarum as revealed by NBD-phallacidin staining.
Cell Biol. Int., 7(8):637–640, 1983.
[13] J-P Rieu, H Delanoe-Ayari, S Takagi, Y Tanaka, and
T Nakagaki. Periodic traction in migrating large amoeba
of Physarum polycephalum. J. R. Soc. Interface, 12(106):
20150099, 2015.
[14] Vladimir A Teplov. Role of mechanics in the appear-
ance of oscillatory instability and standing waves of the
mechanochemical activity in the Physarum polycephalum
plasmodium. J. Phys. D: Appl. Phys., 50(21):213002–22,
2017.
[15] K Matsumoto, T Ueda, and Y Kobatake. Reversal of
thermotaxis with oscillatory stimulation in the plasmod-
ium of Physarum polycephalum. J. Theor. Biol., 131(2):
175–182, 1988.
[16] T Ueda. An intelligent slime mold: A self-organizing sys-
tem of cell shape and information. World Scientific Lec-
ture Notes in Complex Systems: Network of Interacting
Machines. World Scientific Publishing Co., 2005.
[17] S Takagi and T Ueda. Emergence and transitions of dy-
namic patterns of thickness oscillation of the plasmodium
of the true slime mold Physarum polycephalum. Physica
D, 237:420–427, 2008.
[18] S Takagi and T Ueda. Annihilation and creation of rotat-
ing waves by a local light pulse in a protoplasmic droplet
of the Physarum plasmodium. Physica D, 239(11):873–
878, 2010.
[19] Markus Radszuweit, Sergio Alonso, Harald Engel, and
Markus Bar.
Intracellular mechanochemical waves in
an active poroelastic model. Phys. Rev. Lett., 110(13):
138102, 2013.
[20] Sergio Alonso, Ulrike Strachauer, Markus Radszuweit,
Markus Bar, and Marcus J B Hauser. Oscillations and
uniaxial mechanochemical waves in a model of an ac-
tive poroelastic medium: Application to deformation
patterns in protoplasmic droplets of Physarum poly-
cephalum. Physica D, 318-319:58–69, 2016.
[21] T Nakagaki, H Yamada, and T Ueda.
Interaction be-
tween cell shape and contraction pattern in the Physarum
plasmodium. Biophys. Chem., 84(3):195–204, May 2000.
[22] K Alim, G Amselem, F Peaudecerf, M P Brenner, and
A Pringle. Random network peristalsis in Physarum poly-
cephalum organizes fluid flows across an individual. Proc.
Natl. Acad. Sci. U.S.A., 110(33):13306–13311, 2013.
[23] M D. Fricker. personal communication, 2014.
[24] S Kuroda, S Takagi, T Nakagaki, and T Ueda. Allometry
in Physarum plasmodium during free locomotion: size
versus shape, speed and rhythm. J. Exp. Biol., 218(23):
3729–3738, 2015.
[25] T Ueda and Y Kobatake.
Initiation, development and
termination of contraction rhythm in plasmodia of myx-
omycete Physarum polycephalum. J. Theor. Biol., 97(1):
87–93, 1982.
[26] Jonathon Howard, Stephan W Grill, and Justin S Bois.
Turing's next steps: the mechanochemical basis of mor-
phogenesis. Nat. Rev. Mol. Cell Biol., 12(6):392–398,
2011.
4
[27] Peter Gross, K Vijay Kumar, and Stephan W Grill. How
active mechanics and regulatory biochemistry combine to
form patterns in development. Annu. Rev. Biophys., 46
(1):337–356, 2017.
[28] AH Shapiro, MY Jaffrin, and SL Weinberg. Peristaltic
pumping with long wavelengths at low Reynolds number.
J. Fluid Mech., 37:799–825, 1969.
[29] A H Shapiro and M Y Jaffrin. Reflux in peristaltic pump-
ing - is it determined by Eulerian or Lagrangian mean
velocity. J. Appl. Mech., 38(4):1060–1062, 1971.
[30] K P Selverov and H A Stone. Peristaltically driven chan-
nel flows with applications toward micromixing. Phys.
Fluids, 13(7):1837–1859, 2001.
[31] M Li and J G Brasseur. Non-steady peristaltic transport
in finite-length tubes. J. Fluid Mech., 248:129–151, 1993.
[32] M Iima and T Nakagaki. Peristaltic transport and mixing
of cytosol through the whole body of Physarum plasmod-
ium. Math. Med. Biol., 29(3):263–281, 2012.
[33] R Aris. On the dispersion of a solute in a fluid flowing
through a tube. Proc. R. Soc. A, 235:67–77, 1956.
[34] G Taylor. Dispersion of soluble matter in solvent flowing
slowly through a tube. Proc. R. Soc. A, 219:186–203,
1953.
[35] G N Mercer and A J Roberts. A center manifold descrip-
tion of contaminant dispersion in channels with varying
flow properties. SIAM J. Appl. Math., 50(6):1547–1565,
1990.
[36] G N Mercer and A J Roberts. A complete model of shear
dispersion in pipes. Japan J. Indust. Appl. Math., 11(3):
499–521, 1994.
[37] Luke L Heaton, Eduardo L´opez, Philip K Maini, Mark D
Fricker, and Nick S Jones. Advection, diffusion, and de-
livery over a network. Phys. Rev. E, 86(2):021905, 2012.
[38] PG Saffman. A theory of dispersion in a porous medium.
J. Fluid Mech., 6:321–349, 1959.
[39] S Marbach, K Alim, N Andrew, A Pringle, and M P
Brenner.
Pruning to increase Taylor dispersion in
Physarum polycephalum networks. Phys. Rev. Lett., 117
(17):178103, 2016.
[40] K Alim, N Andrew, A Pringle, and M P Brenner. Mech-
anism of signal propagation in Physarum polycephalum.
Proc. Natl. Acad. Sci. U.S.A., 114(20):5136–5141, 2017.
[41] Y Yoshimoto and N Kamiya. Studies on contraction
rhythm of plasmodial strand 3. Role of endoplasmic
streaming in synchronization of local rhythms. Proto-
plasma, 95(1-2):111–121, 1978.
[42] K.E. Samans, I. Hinz, Z. Hejnowicz, and K.E. Wohlfarth-
Bottermann. Phase relation of oscillatory contraction
cycles in Physarum plasmodia: I. A serial infrared regis-
tration device and its application to different plasmodial
stages. J. Interdiscip. Cycle Res., 15(4):241–250, 1984.
[43] U Achenbach and KE Wohlfarth-Bottermann. Synchro-
nization and signal transmission in protoplasmic strands
of Physarum. Planta, 151:574–583, 1981.
[44] D L Taylor, J S Condeelis, P L Moore, and R D Allen.
The contractile basis of amoeboid movement:
I. The
chemical control of motility in isolated cytoplasm. J.
Cell Biol., 59(2):378–394, 1973.
[45] T Lammermann, B L Bader, S J Monkley, T Worbs,
R Wedlich-Soldner, K Hirsch, M Keller, R Forster, D R
Critchley, R Fassler, and M Sixt. Rapid leukocyte mi-
gration by integrin-independent flowing and squeezing.
Nature, 453(7191):51–55, 2008.
[46] G Charras and E Paluch. Blebs lead the way: How to
migrate without lamellipodia. Nat. Rev. Mol. Cell Biol.,
9(9):730–736, 2008.
[47] K Yoshida and T Soldati. Dissection of amoeboid move-
ment into two mechanically distinct modes. J. Cell Sci.,
119:3833–3844, 2006.
[48] K Matsumoto, S Takagi, and T Nakagaki. Locomotive
mechanism of Physarum plasmodia based on spatiotem-
poral analysis of protoplasmic streaming. Biophys. J.,
94:2492–504, 2008.
[49] S Zhang, R D Guy, J C Lasheras, and J C del ´Alamo.
Self-organized mechano-chemical dynamics in amoeboid
locomotion of Physarum fragments. J. Phys. D: Appl.
Phys., 50(20):204004, 2017.
[50] O L Lewis, S Zhang, R D Guy, and J C del Alamo.
Coordination of contractility, adhesion and flow in mi-
grating Physarum amoebae. J. Roy. Soc. Interface, 12
(106):20141359, 2015.
[51] B Rodiek, S Takagi, T Ueda, and M J B Hauser. Patterns
of cell thickness oscillations during directional migration
of Physarum polycephalum. Eur. Biophys. J., 44(5):349–
358, 2015.
[52] W Baumgarten and M J B Hauser. Dynamics of frontal
extension of an amoeboid cell. Europhys. Lett., 108(5):
50010, 2014.
5
[53] A Kamiya and T Togawa. Adaptive regulation of wall
shear-stress to flow change in the canine carotid-artery.
Am. J. Physiol., 239(1):H14–H21, 1980.
[54] C D Murray. The physiological principle of minimum
work: I. The vascular system and the cost of blood vol-
ume. Proc. Natl. Acad. Sci. U.S.A., 12(3):207–214, 1926.
[55] F Corson. Fluctuations and redundancy in optimal trans-
port networks. Phys. Rev. Lett., 104:048703, 2010.
[56] E Katifori, G J Szollosi, and M O Magnasco. Damage and
fluctuations induce loops in optimal transport networks.
Phys. Rev. Lett., 104:048704, 2010.
[57] M Durand. Structure of optimal transport networks sub-
ject to a global constraint. Phys. Rev. Lett., 98(8):088701,
2007.
[58] D Akita, I Kunita, M D Fricker, S Kuroda, KA Sato, and
T Nakagaki. Experimental models for Murray's law. J.
Phys. D: Appl. Phys., 50(2):024001, 2016.
[59] F. J. Meigel and K. Alim. Flow rate of transport network
controls uniform metabolite supply to tissue. Roy. Soc.
Interface, in press, 2018.
[60] S-S Chang, S Tu, Y-H Liu, V Savage, S-P L Hwang, and
M Roper. Optimal occlusion uniformly partitions red
blood cells fluxes within a microvascular network. PLos
Comput. Biol., 13:e1005892, 2017.
|
1607.06836 | 1 | 1607 | 2016-07-22T20:37:37 | Allosteric Regulation by a Critical Membrane | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech",
"q-bio.SC"
] | Many of the processes that underly neural computation are carried out by ion channels embedded in the plasma membrane, a two-dimensional liquid that surrounds all cells. Recent experiments have demonstrated that this membrane is poised close to a liquid-liquid critical point in the Ising universality class. Here we use both exact and stochastic techniques on the lattice Ising model to explore the ramifications of proximity to criticality for proteins that are allosterically coupled to Ising composition modes. Owing to diverging generalized susceptibilities, such a protein's activity becomes strongly influenced by perturbations that influence the two relevant parameters of the critical point, especially the critical temperature. In addition, the protein's kinetics acquire a range of time scales from its surrounding membrane, naturally leading to non-Markovian dynamics. | physics.bio-ph | physics | Allosteric Regulation by a Critical Membrane
Ofer Kimchi,1 Sarah L. Veatch,2 and Benjamin B. Machta3, 1, ∗
1Department of Physics, Princeton University, Princeton, NJ, 08544
2Department of Biophysics, University of Michigan, Ann Arbor, MI. 48109
3Lewis Sigler Institute, Princeton University, Princeton, NJ, 08544
Many of the processes that underly neural computation are carried out by ion channels embedded
in the plasma membrane, a two-dimensional liquid that surrounds all cells. Recent experiments
have demonstrated that this membrane is poised close to a liquid-liquid critical point in the Ising
universality class. Here we use both exact and stochastic techniques on the lattice Ising model to
explore the ramifications of proximity to criticality for proteins that are allosterically coupled to
Ising composition modes. Owing to diverging generalized susceptibilities, such a protein's activity
becomes strongly influenced by perturbations that influence the two relevant parameters of the
critical point, especially the critical temperature. In addition, the protein's kinetics acquire a range
of time scales from its surrounding membrane, naturally leading to non-Markovian dynamics.
PACS numbers: 87.15.kt, 87.15.Ya,87.16.dt
Every cell is surrounded by the plasma membrane,
a two dimensional (2D) liquid composed of lipids and
embedded proteins.
In addition to separating the cell
from its surroundings, the plasma membrane is home
to diverse functional processes. In neurons, membrane-
bound ion channels sense chemical and electrical signals
and control conductance to specific ions, leading to the
complex dynamics that underly neural function. Recent
progress suggests a role for the membrane itself in regu-
lating these processes. The membrane is heterogeneous,
with liquid structures often termed 'rafts' at length scales
of 10 − 100 nm, much larger than the 1 nm scale of
individual lipids [1, 2]. Experiments have suggested a
physical mechanism underlying these structures. Vesi-
cles isolated from a mammalian cell line have membranes
tuned close to a liquid-liquid miscibility critical point [3].
When cooled below their critical temperature, Tc, these
vesicles macroscopically phase separate into two liquid
phases termed liquid ordered (lo) and liquid disordered
(ld) which differ in the partitioning of lipids, proteins and
a fluorescent dye [4].
Previously we have argued that proximity to this criti-
cal point is likely to underly much of the 'raft' heterogene-
ity seen in diverse membrane systems [5] with embedded
proteins subject to long-range critical Casimir forces [6].
More recently we have shown that n-alcohol anesthetics
take membrane derived vesicles away from criticality by
lowering Tc [7]. Despite structural diversity, anesthetics
are known to exert similar effects on diverse ion chan-
nels [8] leading us to speculate that these effects might
arise because anesthetics mimic or interfere with native
regulation of channels by their surrounding membrane.
In support of this, we found that conditions that re-
verse anesthetic effects on ion channels and organisms
also raise critical temperatures in vesicles [9].
Here we explore consequences of thermodynamic criti-
cality for a membrane-bound protein whose internal state
is coupled to the state of its surrounding membrane.
While most ion channels are broadly classified as lig-
and gated or voltage gated, many are also sensitive to
a wide range of modulators including calcium levels, pH,
lipids, and temperature [10].
Ion channel function can
also depend on the 2D solvent properties of the mem-
brane in which they are embedded, both in reconstituted
[11, 12] and in vivo assays [13, 14]. In this manuscript
we develop a simple model for a protein regulated by its
surrounding membrane. We show that when the mem-
brane is held close to a critical point this type of coupling
leads to a qualitatively distinct regime including strong
responses to perturbations which influence Tc, a hierar-
chy of time scales, and non-Markovian dynamics. As we
show, these effects are critical phenomena, arising from
the diverging generalized susceptibilities and correlation
times that emerge near the critical point.
Throughout the manuscript we consider a system con-
sisting of membrane degrees of freedom {s} and a single
hypothetical ion channel (Fig. 1A) with state R and
Hamiltonian
Htot({s}, R) = Hmem({s}) + ER + Hint({s}∂, R)
(1)
where Hmem describes interactions of membrane lipids
with one another, ER measures the (free) energy of state
R without considering membrane interactions, and Hint
describes the interaction between the protein in state R
and the components that it contacts at its boundary,
s ∈ ∂.
In equilibrium, the system consisting of pro-
tein and membrane will be in a given state with prob-
ability determined by an appropriate Boltzmann distri-
bution, P ({s}, R) = e−βHtot({s},R)/Z. We can inte-
grate over membrane degrees of freedom to isolate the
protein whose internal states are occupied according to
P (R) = exp(−βFR)/Z, where
(cid:88)
{s}
exp(−βFR) =
exp (−βHtot({s}, R))
(2)
defines FR, the free energy of state R.
6
1
0
2
l
u
J
2
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
6
3
8
6
0
.
7
0
6
1
:
v
i
X
r
a
2
(See Fig. 1B). These interactions between our protein's
state R and surrounding spins mimic a coupling between
a protein's functional state and the membrane's state as
seen experimentally [11, 13, 14]. They could arise from
hydrophobic mismatch [17, 18], distinct lateral pressure
profiles in the different phases [19], or specific interac-
tions with particular components that partition strongly
into distinct phases [20].
We consider three types of BCs in this manuscript:
fixed BCs in which hi(R) = ±JR, free BCs in which
hi(R) = 0 and Janus BCs in which hi(R) = +JR on one
side of the protein and hi(R) = −JR on the other side.
We first wanted to investigate how a change in the
membrane's properties might influence average channel
activity, as would be measured by a whole cell recording
or other technique probing the response of many channels
together. In the context of our model this means exam-
ining the dependance of P (R) on the details of Hmem.
As this is a static property of our equilibrium system we
can make use of a remarkable algorithm developed for
spin glasses that uses Pfaffian elimination to exactly cal-
culate partition functions to user specified precision on
lattices with arbitrary nearest neighbor couplings Jij in
zero field [21]. We implement our Hamiltonian by set-
ting all interactions Jij between spins to βJ, interactions
between lattice sites internal to the protein to a large
negative value (here −10) and interactions between lat-
tice sites contained in the protein and spins to hi(R) [36]
(see Fig. 1B).
Changes to Hmem will lead to a change in a state's
free energy ∆FR with resulting changes in P (R).
In
this two-state example, only the difference in free en-
ergy changes, ∆Fdif f = ∆F+ − ∆F− will
influence
P (R = ±1) through:
FIG. 1:
(A) Schematic representation of our model. We
consider a protein embedded in a nearly critical membrane
with which it interacts through distinct boundary interactions
in distinct functional states. When the parameters of the
membrane change, the free energies of different protein states
changes, altering function. (B) We probe these effects using
a lattice Ising model in which a single protein is modeled as a
group of sites which transition together, while the remaining
membrane is composed of Ising spins that can take values
si = ±1
As in previous work, we model the membrane as a
square-lattice of spins si = ±1 with the usual Ising
Hamiltonian Hmem = −J(cid:80)(cid:104)i,j(cid:105) sisj, where the sum
runs over all nearest neighbor pairs neither of which
are contained in the protein.
In our model, up/down
spins roughly correspond to more lo/ld partitioning com-
ponents, with a lattice constant which corresponds to
l ∼ 1 − 2 nm of membrane. This scale corresponds both
to the approximate size of a lipid of ∼ .8 nm2[15, 16] and
to measurements of the correlation length near the criti-
cal point of cell-derived vesicles that suggest an analogy
to a lattice model with l ∼ 2 nm [3]. Our model also con-
tains a protein whose internal state can take two values,
R = ±1 which could correspond to open/closed, and a
chemical potential (perhaps modulated by a ligand) that
contributes a term in the Hamiltonian ER = −µR. Cru-
cially we also include an interaction between membrane
i hi(R)si where this
sum is over spins which border the protein, whose bound-
ary conditions (BCs) in state R are determined by hi(R)
and protein of the form Hint = (cid:80)
P (R = +1)
P (R = −1)
→ P (cid:48)(R = +1)
P (cid:48)(R = −1)
exp (−β∆Fdif f )
(3)
Thus exp(−β∆Fdif f ) measures the degree to which a
change in Hmem potentiates the + state (see Fig. 2A).
We estimate the potentiation of function that accom-
panies a 1% change in Tc, comparable both to the natural
variation seen in cell derived vesicles [3], and to the mag-
nitude of the effect seen with physiologically relevant con-
centrations of anesthetics [7]. We examine BCs in which
the R = −1 state has fixed BCs (hi = +∞, here 10 for
numerical purposes) while the R = +1 state has either
free BCs (hi = 0) or Janus BCs. The results are shown in
Fig. 2B-D. In Fig. 2B this potentiation is shown for pro-
teins in which the R = +1 has free BCs, at T = 1.05Tc
for proteins with a range of radii. Larger proteins see
more pronounced potentiation due to the increased sur-
face area of interaction. In Fig. 2C the potentiation is
shown for the same proteins over a range of tempera-
tures. Away from the critical point, this change has only
a small effect on the state occupancy -- the potentiation is
small. However, near the critical point the potentiation
3
the free energy associated with the insertion of an inclu-
sion should have a singular contribution that depends on
inclusion size, BCs and the distance to the critical point:
F s
R = U(rt1/yt, h/tyh/yt , ...)
(4)
where yt = ν = 1 and yh = γ = 15/8 are the scaling
dimension of the fields t and h, r is the radius of the in-
clusion and U is a universal function. Taking a derivative
with respect to t, we expect that
∂F s
dif f
∂t
= rytU(rt1/yt, h/tyh/yt, ...)
(5)
We verify that our measured potentiation can indeed
be collapsed in this manner in Fig. 2D for the fixed/Janus
BCs. However, for the fixed/free particles we find an
additional contribution that arises more directly from the
boundary. This term is proportional to the length of the
boundary and should scale as
F s
R,∂ = r/l U(lt1/yt, h/tyh/yt, ...)
(6)
leading to a similar collapse but with lytt on the x-axis
(see supplement for discussion). While scaling suggests
that coupling to changes in h would diverge more strongly
near criticality, membranes are at fixed magnetization
(composition) in which there is no divergence.
We next looked at the role a critical membrane could
play in shaping the kinetics of ion channel function as
would be measured in single channel voltage clamp ex-
periments. To do this we employ the local Kawasaki
algorithm [22] as we have done previously [5] which im-
plements diffusive 'model B' dynamics [23]. Using these
dynamics a sweep (in which every spin is proposed to ex-
change with its neighbors once) corresponds to roughly
a microsecond of real time since each lattice site corre-
sponds to the size of a lipid and lipids diffuse at a rate
of ∼ 1 µm2/s [5][37]. Our Kawasaki dynamics forbid ex-
change with spins in the protein, but every sweep the
protein changes state with the Metropolis probability to
ensure detailed balance.
While our protein attempts to change its state R dur-
ing every sweep, its apparent kinetics are much slower.
Representative traces of R(τ ) are shown in Fig. 3A af-
ter convolution with a Gaussian of width σ = 103
sweeps [38]. These traces share qualitative features with
real ion channels spanning many time scales [24]: 'flick-
ers' shorter than σ, single 'openings' (which often con-
tain many short flickers), and 'bursts' in which a series
of openings occur close to each other. The time-scale of
these features increases as the critical point is approached
(Fig. 3A). Similar changes in the time-scale of ion chan-
nel dynamics arising from the addition of some anesthet-
ics have been observed experimentally [25]. These could
result from perturbations to the proximity of the mem-
brane to criticality, though they are typically interpreted
FIG. 2: Changes in membrane properties lead to changes
in the conformational equilibrium of a membrane-bound pro-
tein.
(A) In this example P (R = +1) is potentiated by a
change in the critical temperature of the membrane. As µ is
varied (for example through addition of ligand) the protein
transitions from the off (R = −1) to on (R = +1) state. Af-
ter a perturbation has been applied that changes Tc, the free
energy difference between the R = ±1 states has changed by
β∆Fdif f , shifting the curve of P (R = +1) vs µ to the left
(from the black line to the black dashed line). (B) β∆Fdif f
is plotted for a perturbation in which Tc is lowered by 1% for
ion channels of different sizes, in all cases from T = 1.05Tc
and with fixed boundary conditions when R = −1 and free
boundary conditions when R = +1. β∆Fdif f is larger for
larger inclusions, and is ∼ 0.5 for the largest inclusions ex-
amined, comparable to the GABAA channel which is ∼ 6
nm across and which is potentiated by ∼ 50% by analogous
(C) β∆Fdif f is plotted vs t = T − Tc/Tc
treatments [8].
for different sized inclusions (symbols as in B) for the same
fixed-free BCs (red) and for BCs in which R = +1 has Janus
BCs (blue). In each case, the magnitude of β∆Fdif f becomes
much larger closer to the critical point. (D) We verify through
the scaling collapse discussed in the text that this effect can
be understood as a critical phenomenon.
becomes much larger, even for the modest 1% change in
Tc explored here. This change in Tc is comparable to
that seen with addition of clinically relevant concentra-
tions of anesthetic [7, 9]. For the largest proteins sizes
explored here, our observed potentiation is similar to that
observed for the GABAA channel (see Fig. 2 of Ref [8])
whose radius is ∼ 6 nm. While we don't know details of
GABAA's interaction with it's surrounding membrane,
our results suggest that anesthetic effects on membrane
Tc could in principle lead to observed changes in channel
function.
We next wanted to verify that our effect could be un-
derstood as a critical phenomenon. According to scaling,
as arising from direct binding interactions between anes-
thetic and channel.
We quantify these findings by looking at the time auto-
correlation of the simulated protein, χ(τ ) = (cid:104)R(0)R(τ )(cid:105),
for a range of temperatures (Fig. 3B), demonstrating
that correlations in state persist much longer as the crit-
ical point is approached. We also perform simulations
for proteins of different radii all at T = 1.05Tc (Fig. 3C),
demonstrating that larger proteins have slower kinetics
due to the increased area for interaction with surrounding
membrane. These curves quantify the many time-scales
that can be seen qualitatively in Fig. 3A; they are not
closely approximated by single exponentials. Although
the system as a whole is Markovian, when considered in
isolation the protein displays non-Markovian dynamics,
implying that some memory of its history is stored in
the membrane degrees of freedom surrounding it. The
non-Markovianity of several real ion channels and other
membrane-embedded proteins has been well character-
ized and various internal states of these proteins have
been hypothesized [26, 27]; however, our results suggest
that in some cases the history of the protein's state may
be stored in the membrane rather than in internal states
of the protein itself.
We expect that the long time-scales seen in Fig. 3A
are inherited from the surrounding membrane. Near
the Ising critical point the correlation length diverges as
ξ ∝ t−ν, and the correlation time diverges like τcor ∝ ξz
where a dynamic exponent is z = 4 − η = 3.75 for the
Kawasaki dynamics employed here [23]. If the curves of
R(t) reflect properties of these slow critical fluctuations,
then we might expect that autocorrelation functions
in systems with different radii and correlation lengths
χ(τ, r, ξ) might take a universal form:
χ(τ, r, ξ) = rλU (r/ξ, τ /ξz)
(7)
with λ = 2yb = 1 where yb is the scaling dimension of the
boundary operator which is expected to be 1/2 from con-
formal field theory arguments (See [1] and supplement for
a brief discussion). To check this, we performed simula-
tions for the same proteins shown in Fig. 3C but now at
different temperatures chosen so that r/ξ is fixed. While
the bare curves decay over time-scales that differ by a fac-
tor of 102, when we plot χ(τ )/r vs τ /ξz, we see collapse
onto a single curve within stochastic error (Fig. 3D) as
predicted by equation 7.
Conclusion- Regulation by the nearly critical mem-
brane might be very widespread both in ion channels
and for other membrane-bound proteins. There are hints
that the membrane may be playing an important role in
this localization [29] and in some cases in direct regula-
tion [14] of channels. Our model could account for the
sensitivity of many diverse ion channels to chemically di-
verse anesthetics [8] which we have demonstrated lower
the critical temperature of cell-derived vesicles by ∼ 4K,
just over 1% [7]. This explanation is appealing in that
4
FIG. 3:
(A) Simulated ion channel dynamics, R(τ ), have
prominent features at many distinct time-scales, qualitatively
resembling traces from real single-channel recordings [24].
Near Tc, protein state changes occur on a timescale orders of
magnitude larger than that of attempted state changes which
happen once per sweep.
(B) The autocorrelation function
of χ(τ ) is shown at different temperatures, with proximity
to criticality leading to longer lived correlations.
(C) χ(τ )
also depends on channel radius r at a single temperature,
T = 1.05 Tc (colors as above). (D) These long time scales are
critical phenomena. When we plot χ(τ )/r vs τ /ξz for r and
ξ values chosen so that r/ξ is constant, these curves collapse
onto a single universal function as predicted by scaling.
it might also explain why cholesterol depletion, which
changes the membrane's 'magnetization' [30] also effects
many anesthetic-sensitive channels [13]. Our results sug-
gest that such channels should have their partitioning
into small domains modulated by addition of ligand, a
prediction that could be tested using super-resolution
techniques.
While most previous studies, both by ourselves [5, 6]
and others [31] have focused on the role of membrane
thermodynamics in localizing proteins into domains, this
work suggests the same domains could couple more di-
rectly to function. We predict that when a receptor binds
ligand it will change its preference for its surrounding
lipid environment, leading to different interaction part-
ners and an imprint on the state of the local membrane
that will persist even after the ligand has dispersed. We
have highlighted a specific impact of thermodynamic crit-
icality on a single ion channel, and further work will clar-
ify how this mechanism contributes to neural function.
ACKNOWLEDGEMENTS
We thank Bill Bialek, Colin Clement, Jim Sethna and
Ned Wingreen for useful discussions. This work was sup-
ported by NIH R01 GM110052 (SLV) and a Lewis-Sigler
Fellowship (BBM). BBM thanks CK Thomas and AA
Middleton for making their code available on their web-
site (https://aamiddle.expressions.syr.edu/).
∗ Electronic address: [email protected]
[1] Simons K, Toomre D (2000) Lipid rafts and signal trans-
duction. Nature Reviews Molecular Cell Biology 1:31 -- 39.
[2] Dart C (2010) Lipid microdomains and the regulation of
ion channel function. The Journal of Physiology 588:3169 --
3178.
[3] Veatch SL, et al. (2008) Critical fluctuations in plasma
membrane vesicles. Acs Chemical Biology 3:287 -- 293.
[4] Baumgart T, et al. (2007) Large-scale fluid/fluid phase
separation of proteins and lipids in giant plasma mem-
brane vesicles. Proc. Natl. Acad. Sci. 104:3165 -- 3170.
[5] Machta BB, Papanikolaou S, Sethna JP, Veatch SL (2011)
Minimal model of plasma membrane heterogeneity re-
quires coupling cortical actin to criticality. Biophys J
100:1668 -- 77.
[6] Machta BB, Veatch SL, Sethna JP (2012) Critical casimir
forces in cellular membranes. Phys. Rev. Lett. 109:138101.
[7] Gray E, Karslake J, Machta BB, Veatch SL (2013) Liquid
general anesthetics lower critical temperatures in plasma
membrane vesicles. Biophys J 105:2751 -- 9.
[8] Franks NP, Lieb WR (1994) Molecular and cellular mech-
anisms of general anaesthesia. Nature 367:607 -- 14.
5
[17] Andersen OS, Roger E. Koeppe I (2007) Bilayer thickness
and membrane protein function: An energetic perspective.
Annual Review of Biophysics and Biomolecular Structure
36:107 -- 130 PMID: 17263662.
[18] Soubias O, Niu SL, Mitchell DC, Gawrisch K (2008)
Lipid-rhodopsin hydrophobic mismatch alters rhodopsin
helical content. J Am Chem Soc 130:12465 -- 71.
[19] Gruner SM, Shyamsunder E (1991) Is the mechanism of
general anesthesia related to lipid membrane spontaneous
curvature? Ann N Y Acad Sci 625:685 -- 97.
[20] Levitan I, Singh DK, Rosenhouse-Dantsker A (2014)
Cholesterol binding to ion channels. Frontiers in Phys-
iology 5.
[21] Thomas CK, Middleton AA (2013) Numerically exact
correlations and sampling in the two-dimensional ising
spin glass. Phys. Rev. E 87:043303.
[22] Kawasaki K (1972) Phase transitions and critical phe-
(Academic Press,
nomena eds Domb C, Green MS
Waltham, Massachusetts) Vol. 4.
[23] Hohenberg P, Halperin B (1977) Theory of dynamic crit-
ical phenomena. Rev. Mod. Phys. 49:435 -- 479.
[24] Mortensen M, Smart TG (2007) Single-channel recording
of ligand-gated ion channels. Nat. Protocols 2:2826 -- 2841.
[25] Wachtel RE (1995) Relative potencies of volatile anes-
thetics in altering the kinetics of ion channels in bc3h1
cells. Journal of Pharmacology and Experimental Thera-
peutics 274:1355 -- 1361.
[26] Yamashita M, Mori T, Nagata K, Yeh J, Narahashi T
(2005) Isoflurane modulation of neuronal nicotinic acetyl-
choline receptors expressed in human embryonic kidney
cells. Anesthesiology 102:76 -- 84.
[27] Rankin S, Addona G, Kloczewiak M, Bugge B, Miller
K (1997) The cholesterol dependence of activation and
fast desensitization of the nicotinic acetylcholine receptor.
Biophysical Journal 73:2446 -- 2455.
[28] Cardy J
(2004) Boundary Conformal Field Theory.
[9] Machta BB, et al. (2016) Stabilizing membrane domains
ArXiv High Energy Physics - Theory e-prints.
antagonizes n-alcohol anesthesia. In press, Biophys J.
[10] Kinnunen PK (1991) On the principles of functional or-
dering in biological membranes. Chemistry and Physics of
Lipids 57:375 -- 399.
[11] Bristow DR, Martin IL (1987) Solubilisation of the
gamma-aminobutyric acid/benzodiazepine receptor from
rat cerebellum: optimal preservation of the modulatory
responses by natural brain lipids. J Neurochem 49:1386 --
93.
[12] Rankin SE, Addona GH, Kloczewiak MA, Bugge B,
Miller KW (1997) The cholesterol dependence of activa-
tion and fast desensitization of the nicotinic acetylcholine
receptor. Biophysical Journal 73:2446 -- 2455.
[13] Sooksawate T, Simmonds M (2001) Effects of membrane
cholesterol on the sensitivity of the gabaa receptor to gaba
in acutely dissociated rat hippocampal neurones. Neu-
ropharmacology 40:178 -- 184.
[14] Allen JA, Halverson-Tamboli RA, Rasenick MM (2007)
Lipid raft microdomains and neurotransmitter signalling.
Nat Rev Neurosci 8:128 -- 140.
[15] Chiu S, Jakobsson E, Mashl RJ, Scott HL
(2002)
Cholesterol-induced modifications in lipid bilayers: A sim-
ulation study. Biophysical Journal 83:1842 -- 1853.
[16] Alwarawrah M, Dai J, Huang J (2010) A molecular view
of the cholesterol condensing effect in dopc lipid bilay-
ers. The Journal of Physical Chemistry B 114:7516 -- 7523
PMID: 20469902.
[29] Li X, Serwanski DR, Miralles CP, Bahr BA, De Blas AL
(2007) Two pools of triton x-100-insoluble gabaa recep-
tors are present in the brain, one associated to lipid rafts
and another one to the post-synaptic gabaergic complex.
Journal of Neurochemistry 102:1329 -- 1345.
[30] Zhao J, Wu J, Veatch SL (2013) Adhesion stabilizes
robust lipid heterogeneity in supercritical membranes at
physiological temperature. Biophys J 104:825 -- 34.
[31] Katira S, Mandadapu KK, Vaikuntanathan S, Smit B,
Chandler D (2016) Pre-transition effects mediate forces
of assembly between transmembrane proteins.
eLife
5:e13150.
[32] Honerkamp-Smith AR, Machta BB, Keller SL (2012)
Experimental observations of dynamic critical phenomena
in a lipid membrane. Phys. Rev. Lett. 108:265702.
[33] Hamill O, Marty A, Neher E, Sakmann B, Sigworth
F (1981) Improved patch-clamp techniques for high-
resolution current recording from cells and cell-free mem-
brane patches. Pflegers Archiv 391:85 -- 100.
[34] Kasianowicz J, Brandin E, Branton D, Deamer D (1996)
Characterization of individual polynucleotide molecules
using a membrane?channel. Proceedings of the National
Academy of Sciences 93:13770 -- 13773.
ing (Springer Science & Business Media, Berlin).
[35] Sakmann B, Neher E, eds (2009) Single-Channel Record-
[36] This procedure more precisely calculates βF + log 2 since
it integrates over both the correct boundary conditions
6
and those where every hi is flipped.
[37] One caveat is that cells are likely in the regime where
hydrodynamic relaxation [32] competes with diffusive dy-
namics.
[38] Because patch clamp recordings measure times on the or-
der of 50-100 µs [2 -- 4], we plot R(τ ) after convolution with
the Gaussian k(τ ) = e−τ 2/σ2
where σ = 103 (see supple-
ment for bare curves and a discussion). Thus, while our
model has only two states, the recording shows short spikes
reaching levels in between the two, a feature common to
real recordings that presumably has a similar explanation
[2].
(cid:90)
(cid:126)x>r
SUPPLEMENTAL INFORMATION
In supplemental information we first provide a more
detailed argument for the form of the scaling function
needed to capture fixed/free BCs used to make Fig. 2D
of the main text. We then discuss the scaling form of
the dynamic correlation function χ(τ, r, ξ). Finally, we
discuss subtleties in our gaussian blurring procedure used
to make Fig. 3A in the main text.
Scaling of ∂F/∂J for free BCs
In a suitable continuum limit we can write the Hamil-
tonian conditioned on particular boundaries as:
H = J
d(cid:126)x((cid:126)x)
(8)
where ((cid:126)x), the energy operator, is the continuum ver-
sion of ij = si,jsi+1,j. Both here and on the lattice, we
can write the free energy change associated with a change
in Hamiltonian parameters as:
(cid:28)(cid:90)
∂F
∂J
=
(cid:29)
d(cid:126)x((cid:126)x)
(9)
Here we use conformal field theory techniques to ex-
actly calculate the field ((cid:126)r) that surrounds inclusions
with our fixed and free BCs exactly at the critical point
and in the continuum. Although we are unable to carry
out this calculation away from Tc, we can use scaling to
find the form of the relevant function.
At the critical point we make use of exact methods that
we and others have previously used to calculate Casimir
forces between inclusions with particular BCs. Here we
take a different limit, wherein one of the inclusions is
infinitesimal and serves only as a 'test particle' for mea-
suring the strength of the the other inclusion's induced
field ((cid:126)d), which is a function of its BCs.
To calibrate our test particle we first note that ( (cid:126)d) is
a primary scaling field with scaling dimension 1 so that:
(cid:68)
7
+ higher order in d
(cid:69)
UF r−F r(d, r, r) ∼ h2
(r)
((cid:126)0)( (cid:126)d)
(11)
This in conjunction with the exact results allows us to
infer the coupling h(r) ∼ r.
Next we use the same exact results this time with one
protein of radius r and one of infinitesimal radius a0.
At the critical point, conformal invariance allows implies
that their interaction energy can be written as a function
of just one number, x = (d+2a0)(d+2r)
. As a0 becomes
small, x becomes large, even when d << r, allowing
for a measurement of (r) everywhere in space includ-
ing closeby to our inclusion of interest. In this limit, the
scaling form is sufficient, and we find:
ra0
(cid:68)
(cid:69)
((cid:126)d)
=
r
(cid:126)d((cid:126)d + 2r)
(12)
( (cid:126)d)
(cid:69)
(cid:68)
Where the expectation values are conditioned on the
presence of an inclusion with Free BCs of radius r at the
origin, and where d measures the distance from the sur-
face of the protein. We are interested in the integral over
space of
which diverges at both large and short
distances. The short distance divergence we will cut off
by starting our integration at a cutoff motivated by a lat-
tice at l0. Since we are primarily interested in the nearly
critical, rather than critical regime, we assume that the
off-critical energy field takes the following form, which is
certainly true in a regime where ξ > r > l:
( (cid:126)d)
=
r
(cid:126)d( (cid:126)d + 2r)
× U(d/ξ, r/ξ, l/ξ, ...)
(13)
(cid:68)
(cid:69)
where the universal function presumably depends pri-
marily on its first argument (the latter will henceforth
be omitted). Though we do not know the exact form of
U(d/ξ) we expect that it is close to 1 when d/ξ << 1 and
that it goes to 0 when d/ξ >> 1. Our desired integral
can now be written
∂F
∂J
U(d/ξ)
(14)
Now we split this into two terms determined by whether
their contribution diverges as l → 0 or only as ξ → ∞.
2π(r+ (cid:126)d)d (cid:126)d
(cid:90) ∞
(cid:28)(cid:90)
d(d + 2r)
d (cid:126)d((cid:126)d)
(cid:29)
=
=
l0
r
(cid:90) ∞
(cid:90) ∞
l0
πd (cid:126)d r
d
U(d/ξ)
(15)
(16)
(17)
(cid:68)
(0)( (cid:126)d)
(cid:69) ∼ 1
(cid:126)d2
(10)
and
An inclusion with Free BCs in the Ising model will not
couple to the order parameter, and so its leading contri-
butions arise from its coupling to the energy field ((cid:126)x).
When two inclusions of radius r are placed a large dis-
tance d apart, their interactions are closely approximated
as arising from point like insertions of field so that their
interaction energy goes like:
πd (cid:126)d
r
( (cid:126)d + 2r)
U(d/ξ)
l0
In the first we change coordinates to x = d/l0
(cid:90) ∞
πr
1
U ((l/ξ) x)
x
dx
which is of the form r U1(l/ξ) for a (different) universal
function U. For the second term, we can take l → 0 and
instead change variables to x = d/r yielding
(cid:90) ∞
which, for small r1/2Jr, small k and fixed Jr can be writ-
ten as
8
U ((r/ξ) x)
(x + 2)
dx
(18)
χ(τ, r, ξ) = r2ybU(τ /ξz)
(23)
πr
0
which is of the form r U2(r/ξ).
From the above we can write
∂F
∂J
= r U1(l/ξ) + r U2(r/ξ)
(19)
As both terms diverge as the arguments of U1/2 become
large, we generally expect the first to dominate when
except when it is 0 for some reason. This is the case
for the fixed/free BCs as can be seen in Fig. 2D of the
main text. This calculation does not easily extend to
the Janus/fixed case, where (for example) the field ((cid:126)d)
presumably has angular dependance. However, since the
boundary of the janus BCs look locally similar to the
boundary of the fixed BCs, we might expect that the
term with l/ξ dependance cancels, leaving only a term
with r/ξ dependance, as is seen in Fig.2D of the main
paper.
Scaling of dynamics with radius
for
the
argue
scaling
We wish to
form of
χ(τ, r, ξ) = (cid:104)R(0)R(τ )(cid:105) defined in the paper.
First
consider a simpler version where a site magnetic field
adds a term in the Hamiltonian −hs0R. We assume that
at a rate k, R switches or not so as to satisfy detailed
balance so that:
< R(0)R(τ ) >= U(h, kτ, τ /ξz)
(20)
If we assume k is very fast we can ignore the k depen-
dance. Furthermore, h dependance must be even, so that
when h is small we have:
< R(0)R(τ ) >= h2U(τ /ξz) + O(h4)
(21)
where, in fact, this is just (cid:104)R(0)R(τ )(cid:105) = h2 (cid:104)s0(0)s0(τ )(cid:105).
We would like to do a similar calculation here, but now
we must replace h with hef f (r, Jr). We suspect that ours
is best understood as a 'boundary field', whose dimension
is yb = 1/2 [1] so that hef f (r) ∼ r1/2Jr so that
χ(τ, r, ξ) = U(ryb Jr, k τ, τ /ξz)
(22)
This analysis predicts that had we also scaled Jr ∼ r−yb
that we could have found results that collapsed even for
larger proteins for which our relatively large Jrs almost
never flip.
Gaussian Blurring
Patch clamp recordings measure times on the order of
50-100 µs [2 -- 4] while the dynamics of the ion channels
we simulate here occur on timescales of 0.1 µs. There-
fore, in order to simulate the comparitively low temporal
resolution of patch clamp recordings, we plot the state
of the channel as a function of time, R(t), after convolu-
tion with the Gaussian k(τ ) = e−τ 2/σ2
. In Fig. 4 (top),
we show the resulting simulated recordings for a range of
σs spanning several orders of magnitude. Gaussian blur-
ring leads to the emergence of a timescale of ion channel
dynamics on the order of tens of ms, and this result is
robust to changes in σ. This feature appears to be a non-
Markovian phenomenon: a similar analysis performed on
an ion channel simulated at high temperature far from
criticality, for which the channel's dynamics are Marko-
vian, does not lead to the emergence of such timescales
(Fig. 4, bottom).
∗ Electronic address: [email protected]
[1] Cardy J (2004) Boundary Conformal Field Theory. ArXiv
High Energy Physics - Theory e-prints.
[2] Hamill O, Marty A, Neher E, Sakmann B, Sigworth
F (1981) Improved patch-clamp techniques for high-
resolution current recording from cells and cell-free mem-
brane patches. Pflegers Archiv 391:85 -- 100.
[3] Kasianowicz J, Brandin E, Branton D, Deamer D (1996)
Characterization of individual polynucleotide molecules
using a membrane?channel. Proceedings of the National
Academy of Sciences 93:13770 -- 13773.
[4] Sakmann B, Neher E, eds (2009) Single-Channel Record-
ing (Springer Science & Business Media, Berlin).
9
FIG. 4: Results Are Robust to Resolution of Gaussian
Blurring Top: The simulated protein attempted to flip every
sweep (corresponding to 0.1 µs), and without any blurring,
the simulated ion channel recording includes state changes on
the corresponding timescale (top left). However, when the
low temporal resolution of standard patch-clamp techniques
is accounted for through convolution of the data with the
Gaussian k(t) = e−t2/σ2
as in Fig. 3A, a timescale for protein
dynamics on the order of tens of ms can be seen. This result
is not changed when the resolution of the Gaussian blurring,
σ, is varied, even by orders of magnitude. Bottom: No such
timescale emerges in the Markovian case.
|
1608.06918 | 1 | 1608 | 2016-08-19T14:51:33 | Growth and Scaling during Development and Regeneration | [
"physics.bio-ph",
"q-bio.TO"
] | Life presents fascinating examples of self-organization and emergent phenomena. In multi-cellular organisms, a multitude of cells interact to form and maintain highly complex body plans of well-defined size. In this thesis, we investigate theoretical feedback mechanisms for both self-organized body plan patterning and size control.
The thesis is inspired by the astonishing scaling and regeneration abilities of flatworms. These worms can perfectly regrow their entire body plan even from tiny amputation fragments like the tip of the tail. Moreover, they can grow and actively de-grow by more than a factor of 40 in length depending on feeding conditions. These capabilities prompt for remarkable physical mechanisms of self-organized pattern formation and scaling.
First, we explore the basic principles and challenges of pattern scaling in mechanisms previously proposed to describe biological pattern formation. Next, we present a novel class of patterning mechanisms yielding entirely self-organized and self-scaling patterns. This framework captures essential features of body plan regeneration and scaling in flatworms. Further, we analyze shape and motility of flatworms. By applying principal component analysis, we characterize shape dynamics during different motility modes and also identify shape variations between different flatworm species. Finally, we investigate the metabolic control of cell turnover and growth. We identify three mechanisms of metabolic energy storage; theoretical descriptions thereof can explain the measured organism growth by rules on the cellular scale.
In a close collaboration with experimental biologists, we combine minimal theoretical descriptions with state-of-the-art experiments and data analysis. This allows us to identify generic principles of scalable body plan patterning and growth control in flatworms. | physics.bio-ph | physics |
Max-Planck-Institut fur
Physik komplexer Systeme
Growth and Scaling during
Development and Regeneration
Dissertation
zur Erlangung des akademischen Grades
Doctor rerum naturalium
Institut fur Theoretische Physik
Fakultat fur Mathematik und Naturwissenschaften
Technische Universitat Dresden
vorgelegt von
Steffen Werner
geboren am 3. September 1984 in Bad Mergentheim
Dresden, 2016
Eingereicht am 04. Dezember 2015
Promotionskommission:
Prof. Dr. D. Inosov (Vorsitz)
Prof. Dr. F. Julicher (Gutachter)
Prof. Dr. St. Grill (Gutachter)
Prof. Dr. N. Barkai (Gutachter)
Prof. Dr. G. Pospiech
Verteidigt am 17. Juni 2016
Abstract in English
Life presents fascinating examples of self-organization and emergent phenomena. In
multi-cellular organisms, a multitude of cells interact to form and maintain highly complex body
plans. This requires reliable communication between cells on various length scales. First, there
has to be the right number of cells to preserve the integrity of the body and its size. Second,
there have to be the right types of cells at the right positions to result in a functional body
layout. In this thesis, we investigate theoretical feedback mechanisms for both self-organized
body plan patterning and size control.
The thesis is inspired by the astonishing scaling and regeneration abilities of flat-
worms. These worms can perfectly regrow their entire body plan even from tiny amputation
fragments like the tip of the tail. Moreover, they can grow and actively de-grow by more than a
factor of 40 in length depending on feeding conditions, scaling up and down all body parts while
maintaining their functionality. These capabilities prompt for remarkable physical mechanisms
of pattern formation.
First, we explore pattern scaling in mechanisms previously proposed to describe
biological pattern formation. We systematically extract requirements for scaling and high-
light the limitations of these previous models in their ability to account for growth and regene-
ration in flatworms. In particular, we discuss a prominent model for the spontaneous formation
of biological patterns introduced by Alan Turing. We characterize the hierarchy of steady states
of such a Turing mechanism and demonstrate that Turing patterns do not naturally scale.
Second, we present a novel class of patterning mechanisms yielding entirely self-
organized and self-scaling patterns. Our framework combines a Turing system with our
derived principles of pattern scaling and thus captures essential features of body plan regenera-
tion and scaling in flatworms. We deduce general signatures of pattern scaling using dynamical
systems theory. These signatures are discussed in the context of experimental data.
Next, we analyze shape and motility of flatworms. By monitoring worm motility, we can
identify movement phenotypes upon gene knockout, reporting on patterning defects in the lo-
comotory system. Furthermore, we adapt shape mode analysis to study 2D body deformations
of wildtype worms, which enables us to characterize two main motility modes: a smooth gliding
mode due to the beating of their cilia and an inchworming behavior based on muscle contrac-
tions. Additionally, we apply this technique to investigate shape variations between different
flatworm species. With this approach, we aim at relating form and function in flatworms.
Finally, we investigate the metabolic control of cell turnover and growth. We estab-
lish a protocol for accurate measurements of growth dynamics in flatworms. We discern three
mechanisms of metabolic energy storage; theoretical descriptions thereof can explain the ob-
served organism growth by rules on the cellular scale. From this, we derive specific predictions
to be tested in future experiments.
In a close collaboration with experimental biologists, we combine minimal theoret-
ical descriptions with state-of-the-art experiments and data analysis. This allows us
to identify generic principles of scalable body plan patterning and growth control in flatworms.
i
Zusammenfassung auf Deutsch
Die belebte Natur bietet uns zahlreiche faszinierende Beispiele fur die Phanomene
von Selbstorganisation und Emergenz. In Vielzellern interagieren Millionen von Zellen
miteinander und sind dadurch in der Lage komplexe Korperformen auszubilden und zu unter-
halten. Dies verlangt noch einer zuverlassigen Kommunikation zwischen den Zellen auf ver-
schiedenen Langenskalen. Einerseits ist stets eine bestimmte Zellanzahl erforderlich, sodass der
Korper intakt bleibt und seine Grosse erhalt. Anderseits muss fur einen funktionstuchtigen
Korper aber auch der richtige Zelltyp an der richtigen Stelle zu finden sein. In der vorliegenden
Dissertation untersuchen wir beide Aspekte, die Kontrolle von Wachstum sowie die selbstor-
ganisierte Ausbildung des Korperbaus.
Die Dissertation ist inspiriert von den erstaunlichen Skalierungs- und Regene-
rationsfahigkeiten von Plattwurmern. Diese Wurmer konnen ihren Korper selbst aus
winzigen abgetrennten Fragmenten -- wie etwa der Schwanzspitze -- komplett regenerieren.
Daruberhinaus konnen sie auch, je nach Futterungsbedingung, um mehr als das 40fache in der
Lange wachsen oder schrumpfen und passen dabei alle Korperteile entsprechend an, wobei deren
Funktionalitat erhalten bleibt. Diese Fahigkeiten verlangen nach bemerkenswerten physikali-
schen Musterbildungsmechanismen.
Zunachst untersuchen wir das Skalierungsverhalten von fruheren Ansatzen zur
Beschreibung biologischer Musterbildung. Wir leiten daraus Voraussetzung fur das
Skalieren ab und zeigen auf, dass die bekannten Modelle nur begrenzt auf Wachstum und
Regeneration von Plattwurmern angewendet werden konnen.
Insbesondere diskutieren wir
ein wichtiges Modell fur die spontane Entstehung von biologischen Strukturen, das von Alan
Turing vorgeschlagen wurde. Wir charakterisieren die Hierarchie von stationaren Zustanden
solcher Turing Mechanismen und veranschaulichen, dass diese Turingmuster nicht ohne weiteres
skalieren.
Daraufhin prasentieren wir eine neuartige Klasse von Musterbildungsmechanis-
men, die vollstandig selbstorgansierte und selbstskalierende Muster erzeugen. Unser
Ansatz vereint ein Turing System mit den zuvor hergeleiteten Prinzipien fur das Skalieren von
Mustern und beschreibt dadurch wesentliche Aspekte der Regeneration und Skalierung von
Plattwurmern. Mit Hilfe der Theorie dynamischer Systeme leiten wir allgemeine Merkmale von
skalierenden Mustern ab, die wir im Hinblick auf experimentelle Daten diskutieren.
Als nachstes analysieren wir Form und Fortbewegung der Wurmer. Die Auswer-
tung des Bewegungsverhaltens, nachdem einzelne Gene ausgeschaltet wurden, ermoglicht Ruck-
schlussse auf die Bedeutung dieser Gene fur den Bewegungsapparat. Daruber hinaus wenden
wir eine Hauptkomponentenanalyse auf die Verformungen des zweidimensionalen Wurmkorpers
wahrend der naturlichen Fortbewegung an. Damit sind wir in der Lage, zwei wichtige Fortbe-
wegungsstrategien der Wurmer zu charakterisieren: eine durch den Zilienschlag angetriebene
gleichmassige Gleitbewegung und eine raupenartige Bewegung, die auf Muskelkontraktionen
beruht. Zusatzlich wenden wir diese Analysetechnik auch an, um Unterschiede in der Gestalt
von verschiedenen Plattwurmarten zu untersuchen. Grundsatzlich zielen alle diese Ansatze da-
rauf ab, das Aussehen der Plattwurmer mit den damit verbundenen Funktionen verschiedener
Korperteile in Beziehung zu setzen.
ii
Schlussendlich erforschen wir den Einfluss des Stoffwechsels auf den Zellaustausch
und das Wachstum. Dazu etablieren wir Messungen der Wachstumsdynamik in Plattwurmern.
Wir unterscheiden drei Mechanismen fur das Speichern von Stoffwechselenergie, deren theoreti-
sche Beschreibung es uns ermoglicht, das beobachtete makroskopische Wachstum des Organis-
mus mit dem Verhalten der einzelnen Zellen zu erklaren. Basierend darauf leiten wir Vorher-
sagen ab, die nun experimentell getestet werden.
In enger Zusammenarbeit mit Kollegen aus der experimentellen Biologie fuhren wir
minimale theoretische Beschreibungen mit modernsten Experimenten und Analyse-
techniken zusammen. Dadurch sind wir in der Lage, Grundlagen sowohl der skalierbaren
Ausbildung des Korperbaus als auch der Wachstumskontrolle bei Plattwurmern herauszuar-
beiten.
iii
Acknowledgements
I would like to acknowledge all the people who supported me during my
PhD. My special thanks go to:
My supervisors Frank Julicher und Benjamin Friedrich for their great
support and assistance and for teaching me their inspiring way of doing
science. Thanks to Frank Julicher for his enthusiasm about my projects, for
always making sure to be available for insightful discussions and advice and
generally for all the stimulating input. Thanks to Benjamin Friedrich for
his valuable and highly appreciated guidance, for encouraging me to pursue
my own ideas and for always being there to help structuring my work, to
make sure I stay on track and to provide new food for thoughts.
Jochen Rink for being the driving force in this exciting and fruitful colla-
boration, for his open-mindedness towards our theoretical ideas and for
sharing his visionary thoughts on the flatworm project with us. Thanks
also for creating such an enjoyable and inspiring atmosphere in the lab.
Tom Stuckemann and Albert Thommen not only for working hard to
produce some amazing data but also for sharing this data with me, which
triggered major parts of this thesis. Thanks for all the motivating discus-
sions and the productive team spirit.
Nicole Alt, Johanna Richter and many further student helpers for
putting a lot of effort and dedication in taming the beasts under the macro-
scope, generating beautiful movies for me to analyze.
Sarah Mansour, Olga Frank, James Cleland and Shang-Yun Liu for
the exciting times we are having discussing science and beyond.
All the other members of the Rink lab for creating such a fun atmos-
phere to work in.
Lutz Brusch and Michael Kucken for many fruitful discussions during
the theory gatherings.
iv
Daniel Aguilar-Hidalgo for being a motivating sparring partner with
respect to the scaling project and for the great time we had developing
theories.
The summer interns Manuel Beir´an-Amigo and Yihui Quek for the
intruiging questions they asked and for their enthusiasm about the flatworm
project creating an extra thrust pushing it forward.
Ulrike Burkert and many other members of the administration
and the IT department for making everything run so smoothly and for
the quick and non-bureaucratic help with no matter what question.
Vivien Scherr and her colleagues from the library for the great efforts
to track down any exotic piece of literature I was asking for.
My colleagues and friends for contributing in many ways to provide a
great working environment and to make my time at the PKS really enjoy-
able. Thanks for all the discussions and the advice about science and life in
general, all the coffee breaks, kicker sessions, whisky tastings, tango lessons,
movie nights, sciency pub outings ...
My parents and my brother Jochen for their enduring support during
my whole life and for the great feeling that I always can rely on them.
Ruth for all her patience and support, for her understanding of my late
night working sessions and for giving me great backup during the final
stretch as well as for the proofreading of my thesis.
I also gratefully acknowledge the funding by the Max Planck Society and
the German Federal Ministry of Education and Research (BMBF).
v
Contents
1. Introduction
1.1. Development, growth and regeneration . . . . . . . . . . . . . . . . . . .
1.1.1. From cells to tissues to organisms
. . . . . . . . . . . . . . . . .
1.1.2. Cellular communication and chemical signals . . . . . . . . . . .
1.1.3. From signals to genes and back . . . . . . . . . . . . . . . . . . .
1.1.4. Gene expression can be modified in experiments
. . . . . . . . .
1.2. Flatworms as a model organism for scaling, growth and regeneration . .
1.3. Theories of body plan patterning by morphogens . . . . . . . . . . . . .
1.4. Turing mechanism yields self-organized patterns
. . . . . . . . . . . . .
1.5. Open questions in growth control and scalable body patterning . . . . .
1.6. Organization of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . .
2. Scaling in morphogen systems
2.1. Scaling of biological patterns
. . . . . . . . . . . . . . . . . . . . . . . .
2.2. Morphogen dynamics and concentration profiles . . . . . . . . . . . . . .
2.3. Scaling of concentration profiles with system size . . . . . . . . . . . . .
2.4. Scaling of morphogen profiles in pre-patterned systems . . . . . . . . . .
2.4.1. The concept of an expander as a chemical size reporter
. . . . .
2.4.2. A simple mechanism of gradient scaling:
1
1
2
4
6
8
8
13
16
19
21
22
22
23
24
27
27
The expander-dilution model
. . . . . . . . . . . . . . . . . . . .
28
2.4.3. Perfect shape scaling vs. approximate scaling:
The expansion-repression model
. . . . . . . . . . . . . . . . . .
2.4.4. Other feedback schemes for self-organized gradient scaling . . . .
2.4.5. Conclusions from our analysis of expander models
. . . . . . . .
2.5. Revisiting the absence of scaling in Turing patterns . . . . . . . . . . . .
2.5.1. A hierarchy of steady state solutions . . . . . . . . . . . . . . . .
2.5.2. Higher order patterns form in larger systems
. . . . . . . . . . .
2.6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
29
32
32
34
35
38
41
vi
Contents
3. Scaling and regeneration of self-organized patterns
3.1. Self-scaling and self-organization . . . . . . . . . . . . . . . . . . . . . .
3.2. A minimal model for self-organized pattern scaling . . . . . . . . . . . .
3.3. Numeric solution shows scaling and pattern regeneration . . . . . . . . .
3.4. Dynamical systems analysis of scaling and regeneration . . . . . . . . .
42
42
43
44
44
3.5. As single source pattern is attractive for a large region of the phase space 50
3.6. Structural robustness for pattern scaling . . . . . . . . . . . . . . . . . .
3.7. Signatures of self-scaling patterns . . . . . . . . . . . . . . . . . . . . . .
3.8. Comparison to experiments in flatworms . . . . . . . . . . . . . . . . . .
3.9. Summary and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. Flatworm shape dynamics and motility
4.1. Modes, movement and morphology . . . . . . . . . . . . . . . . . . . . .
4.2. Worm motility reports on patterning defects upon gene knockout . . . .
4.3. Shape mode analysis of 2D worm outlines . . . . . . . . . . . . . . . . .
4.4. Shape dynamics during crawling and inchworming . . . . . . . . . . . .
4.4.1. A bending mode and two width-changing modes
. . . . . . . . .
4.4.2. The second and third modes characterize inch-worming . . . . .
4.5. Discriminating flatworm species by shape . . . . . . . . . . . . . . . . .
4.6. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5. Quantitative study of flatworm growth and cell turnover
5.1. Homeostasis is a dynamic steady state . . . . . . . . . . . . . . . . . . .
5.2. Size-dependent growth and degrowth dynamics in flatworms . . . . . . .
5.2.1. Allometric scaling laws . . . . . . . . . . . . . . . . . . . . . . . .
5.2.2. Characterizing the immediate growth response upon feeding . . .
5.2.3. Small worms grow and degrow faster than large worms . . . . . .
5.3. Theoretical descriptions of cell turnover dynamics and energy flux . . .
5.3.1. Model 1: Dynamic energy storage
. . . . . . . . . . . . . . . . .
5.3.2. Model 2: Fixed proportion energy storage . . . . . . . . . . . . .
5.3.3. Model 3: Size-dependent energy storage . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . .
5.3.4. Discussion of the turnover models
5.4. Control logic for cell turnover and growth . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . .
5.4.1. Measuring cell turnover on various scales
5.4.2. Analyzing turnover of the epidermis as an example tissue . . . .
5.5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6. Summary and outlook
51
53
54
56
58
58
60
63
65
65
66
66
68
70
70
72
72
77
79
83
85
87
89
89
91
91
92
97
98
vii
Contents
A. Reaction-diffusion systems: fixed points and scaling
102
A.1. Morphogen dynamics with linear degradation . . . . . . . . . . . . . . . 102
A.1.1. Reaction, diffusion, advection and dilution . . . . . . . . . . . . . 102
A.1.2. Steady state solution neglecting tissue growth . . . . . . . . . . . 102
A.1.3. Relaxation to the steady state
. . . . . . . . . . . . . . . . . . . 103
A.1.4. Steady state without morphogen degradation in the source
. . . 104
A.2. Gradient scaling with expander . . . . . . . . . . . . . . . . . . . . . . . 104
A.2.1. On the scaling with an autonomously controlled expander . . . . 104
A.2.2. Scaling by expander feedback with a switch-like production . . . 105
A.2.3. Scaling by expander feedback with a graded production . . . . . 106
A.3. Linear stability analysis of a Turing system . . . . . . . . . . . . . . . . 107
A.3.1. Eigenvalues of the linearized reaction-diffusion matrix . . . . . . 107
A.3.2. The principle of local activation and lateral inhibition . . . . . . 109
A.4. Motivation for the Hill function . . . . . . . . . . . . . . . . . . . . . . . 110
A.5. Homogeneous steady state of our Turing system . . . . . . . . . . . . . . 112
A.6. Inhomogenous steady states of our Turing system . . . . . . . . . . . . . 114
A.6.1. First order steady state solution . . . . . . . . . . . . . . . . . . 114
A.6.2. Source size of the first order steady state
. . . . . . . . . . . . . 115
A.6.3. Hierarchy of higher order steady states . . . . . . . . . . . . . . . 117
A.6.4. Stability of the inhomogeneous steady state of our Turing system 117
A.7. On our scalable Turing system . . . . . . . . . . . . . . . . . . . . . . . 120
A.7.1. A homogeneous dynamic state for low expander levels . . . . . . 120
A.7.2. Generalized scalable Turing system . . . . . . . . . . . . . . . . . 121
A.7.3. Scaling of downstream targets with a constant amplitude . . . . 121
A.7.4. On knockout experiments in scalable Turing systems . . . . . . . 122
B. On the numerical solution of reaction-diffusion equations
125
B.1. Euler method and Courant criterion . . . . . . . . . . . . . . . . . . . . 125
B.2. Algorithmic speed-up using a convolution with a Gauss kernel . . . . . . 126
C. Worm handling and measurements of size and shape
129
C.1. Worm handling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
C.2. Image acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
C.3. Extracting size and shape from worm movies
. . . . . . . . . . . . . . . 130
D. Shape reconstruction and worm bending
132
D.1. On the reconstruction of a closed worm outline . . . . . . . . . . . . . . 132
D.2. On the reconstruction of head shapes . . . . . . . . . . . . . . . . . . . . 132
viii
Contents
E. On growth and cell turnover
133
. . . . . . . . . . . . . . . . . . . . . . 133
E.1. Signatures of aging in flatworms
E.2. Additional size measurements and growth dynamics
. . . . . . . . . . . 134
E.3. Measuring cell cycle times . . . . . . . . . . . . . . . . . . . . . . . . . . 137
E.4. Measuring cell turnover dynamics on the organism level
. . . . . . . . . 138
E.4.1. Measurement of C14 reveals dynamics of neurogenesis in humans 138
E.4.2. Adapting the C14 technique to cell turnover in flatworms
. . . . 140
E.4.3. Label dynamics after a single feeding event
. . . . . . . . . . . . 142
E.4.4. Challenges and advantages of the application to flatworms . . . . 144
List of Figures
List of Variables
References
145
147
150
ix
1. Introduction
1.1. Development, growth and regeneration
The world around us is populated by a great variety of organisms of very different
shapes, sizes and levels of complexity. Many of the most complex organisms, including
humans, develop from a single fertilized egg cell,
see Fig. 1.1(a). The egg divides multiple times
to give rise to the many cells that form the dif-
ferent tissues of the adult organism (68, 244).
This embryonic development results in a well-
defined body plan of the organism, which eventu-
ally can reproduce again. One important aspect
"Cell and tissue, shell and bone, leaf and flower,
... Their problems of form are in the first in-
stance mathematical problems, their problems of
growth are essentially physical problems, and the
morphologist is, ipso facto, a student of physical
science." -- D'Arcy W. Thompson, On Growth
and Form, 1945 (217)
of development is growth, i.e. the increase in organism size. The growth at the scale of
the organism follows from processes at the cellular scale: (i) an increase in cell number
by cell division, (ii) an increase in cell size by cellular growth and (iii) an increase in
the extra-cellular material (244).
During most of its lifetime, an organism maintains shape and function of its body,
despite the fact that cells continuously become damaged and get lost (68, 158, 244).
This homeostasis requires the sustained addition of new cells by cell division as well
as mechanisms of controlled cell death such as apoptosis. Importantly, the turnover
processes have to be well orchestrated at the cell, tissue and organism level. Imperfect
homeostasis results in aging of the organism (158).
Furthermore, many organisms can regenerate after injury to some extent, see Fig. 1.1(b)
(68, 184, 244). Regeneration refers to de-novo formation of large parts of tissues and
organs that have been damaged or lost. In contrast to embryonic development, which
comprises a fixed sequence of morphogenetic events starting from the fertilized egg as
a well-defined initial condition, the starting point of regeneration strongly depends on
the injury and is thus variable. It is a major question to what extent both processes
are guided by the same principles and depend on the same mechanisms (184).
The ability to reproduce and the permanent struggle against decay are important cha-
racteristics of life in general (192). We are only beginning to understand the respective
1
1. INTRODUCTION
Figure 1.1.: (a) Human development starts from a single fertilized egg cell, from
which the body plan emerges and grows to its final size.
(b) Many multi-cellular
organisms like lizards and salamanders can regenerate major parts of their body.
processes of development, growth and regeneration in simple model systems with the
help of modern molecular biology. In this thesis, we combine minimal theoretical de-
scriptions with the analysis of biological data in flatworms as model systems in order
to extract fundamental physical principles for body plan patterning and growth control
in multicellular organisms.
1.1.1. From cells to tissues to organisms
The question of how an organism forms has puzzled natural philosophers and researchers
for more than two millennia. It was Aristotle who performed one of the earliest do-
cumented experiments in developmental biology in 350 BC (68, 216, 244). He opened
chicken eggs at various time intervals after fertilization and observed that the embryo
gradually resembles a chicken. The gradual formation of an organism, called epige-
nesis, was confronted with the alternative hypothesis of preformation (68, 244). The
latter assumes a completely pre-formed miniature body, which then only grows. De-
spite the work of Aristotle, the theory of preformation was still prevalent in Europe
until the 18th century and is embodied in the idea of the "homunculus", the tiny version
2
1.1 Development, growth and regeneration
of a person encapsulated in the sperm cell (68, 228, 244). Later a related discussion
arose about the concepts of pre-encoded and emergent complexity, as will be mentioned
below (188).
The basic building blocks of higher order organisms are the cells, which come at very
different shapes and sizes (68, 244). For example, a muscle cell and a blood cell have a
completely different appearance and very different properties but they both originate
from a single egg cell, which has divided many times to give rise to all the cells of
the body (68, 244). The cells become committed to fulfill specific tasks and change
their properties during the process of differentiation. Uncommitted cells that can
differentiate into other cell types are called stem cells. Often, there are also tissue-
specific stem cells that are already partially committed and can only turn into a subset
of cell types. In some organisms, differentiated cells are able to de-differentiate into less
committed cells (15, 214).
During embryonic development and regeneration, cells differentiate and organize in a
position dependent manner to form a well-defined body plan. Most modern animals
are Bilaterians, characterized by three main body axis, see Fig. 1.2(a): the anterior-
posterior axis from tail to head, the dorso-ventral axis from the front to the backside
and a mirror symmetric medio-lateral axis (141). Yet, how is the cellular behavior
choreographed with respect to this internal coordinate system?
With the advancements in light microscopy and the ability to observe microscopic
structures, biologists could address this question and perform experiments, in which
they selectively perturbed a specific part of an organism in order to reveal its functions
in body plan patterning (42, 68, 108, 125, 128, 244). As a result, biological research
changed from a descriptive to an experimental science. Here, we will highlight three
early experiments by Chabry, Driesch and Morgan to discuss important concepts of
morphogenesis.
Chabry selectively killed individual cells in the early embryo of the marine invertebrate
Tunicate after the first or second cell division with a needle.
In consequence, only
certain parts of the organism developed, depending on which cells he had destroyed
(68, 188). The results were later confirmed by completely removing the two muscle
precursor cells of the 8-cell embryo (68, 241). These separated cells became muscle
cells by themselves, while the remaining embryo was lacking the muscles. Driesch
performed a similar experiment in sea-urchins but obtained completely opposite results
(42). He separated the two cells after the first round of cell division and observed that
a single cell can develop into a perfectly patterned organism, just of a smaller size.
3
1. INTRODUCTION
Figure 1.2.: (a) The bilaterian body plan, for example of a frog, is characterized by
three perpendicular axes in anterior-posterior (AP), dorso-ventral (DV) and medio-
lateral (ML) direction. (b) Two concepts of morphogenesis: mosaic development based
on pre-encoded structures (exemplified by purple cells forming the head), self-organized
formation of a body plan as an emergent phenomenon based on mutual interactions
between cells.
At the same time, Morgan was performing various regeneration experiments especially
with the freshwater polyp Hydra and flatworms, and he reported that these animals
could re-grow perfectly patterned heads, tails and other body parts after amputation
(128).
These observations can be discussed in the light of two fundamental concepts of mor-
phogenesis, see Fig. 1.2(b) (68, 188): First, the theory of self-differentiation (or mosaic
development) builds on the idea of a pre-encoded (hidden) complexity in the early par-
titioning of the tissue that then only enfolds. Second, the converse theory (sometimes
called conditional specification) considers complexity of the body plan as an emergent
phenomenon by the interaction of different parts. Today, we begin to appreciate that
embryonic development combines both paradigms. The second one has the appeal to
account for regeneration in a natural fashion and will be studied in this thesis in the
context of self-organized body plan patterning.
1.1.2. Cellular communication and chemical signals
At the end of the 19th and the beginning of the 20th century, the existence of "forma-
tive substances" was postulated to control cell fate during development and regenera-
tion (128, 129, 174, 188). It was proposed that these substances are found in graded
abundance originating from the animal poles as the organizing centers and that they
determine polarity and growth of an organism by controlling the cellular behavior
(128, 129, 174). Related concepts were built on a "physiological gradient", inspired
for example by the fact that the regenerative capability in some flatworm species varies
along the body axis from head to tail (36).
4
AMVDPLL(a)(b)pre-encodedself-organized??body planbody plan1.1 Development, growth and regeneration
These considerations lead to the notion of morphogens as specific signaling molecules
that are secreted in distinct source regions and spread in the tissue. The morphogen
concentrations provide chemical cues that control division and differentiation of cells
(68, 79, 138, 244). The term was originally coined by Turing, who proposed a purely
theoretical mechanism for the spontaneous emergence of chemical patterns as a tem-
plate for the body plan layout (221). As a complementary theoretical approach, it
was discussed how a pre-existing organizing region can instruct tissue patterning by
secretion of morphogens (174, 243). It was proposed that graded concentration profiles
provide cells with the information about their position within the tissue. We will dis-
cuss a well-known illustration of this idea, the French flag model, in Section 1.3. Next,
we provide biological examples for organizing regions and concentration gradients of
signaling molecules.
Early experiments by Mangold and Spemann found evidence for an organizing region
that instructs body plan patterning in the embryo of the frog Xenopus laevis (204).
When transplanting the now so-called Spemann organizer into another frog embryo,
the latter developed a second perfectly patterned head. Furthermore, experimental
evidence for "organizing substances" at the animal poles was found in leaf hoppers
(100, 186, 187). After splitting the embryo in a head and a tail fragment, in most cases
neither part developed normally. Yet, if substances from the tail tip were moved to
the head fragment, the head fragment developed into a complete embryo. Interestingly,
also the tail fragment developed further, indicating a concentration-dependent effect of
these tail substances.
Pioneering work by Nusslein-Volhard and colleagues could identify the protein Bicoid
as a signaling molecule in the embryo of the fruit fly Drosophila melanogaster and
demonstrated its instructive role in tissue patterning, see Fig. 1.3(a) (45, 52, 56). They
also visualized its graded concentration profile decreasing from the anterior tip, which
could be fitted by an exponential function (43, 44). Additionally, they showed that
Bicoid influences cellular behavior in a concentration-dependent manner (44).
There are several key signaling systems for patterning and growth control that are
widespread across the animal kingdom. Prominent examples belong to the Transform-
ing growth factor β (TGF-β) superfamily and to the Wnt family. TGF-β proteins can
be found in a wide range of organisms from simple worms to mammals. They control
growth, patterning, tissue homeostasis and even the immune system (88). In this thesis,
we will encounter four examples of these proteins: Activin in the clawed frog Xenopus
(72, 75, 80) and Decapentaplegic (Dpp) in the fruit fly (5, 109, 205, 234), see Fig. 1.3(b),
5
1. INTRODUCTION
Figure 1.3.: (a) Bicoid protein gradient (black) in the embryo of the fruit fly
Drosophila melanogaster at different stages of development (anterior side at the left,
modified with permission from (43), scale bar and approximate timing added by the
author) (b) Decapentaplegic protein in the imaginal wing disc of the fruit fly labeled
by GFP and quantification by GFP fluorescence (green) and GFP immunostaining
(red) (modified with permission from (109)).
as well as Bone morphogenetic protein (Bmp) and Anti-dorsalizing morphogenic pro-
tein (Admp) in flatworms (4, 61). Also Wnt family members are found in organisms
from invertebrates to humans (9, 40, 107, 141, 144, 164). The name is a portmanteau
of Wingless (the corresponding protein in the fruit fly), and Integration 1 (the homolog
originally identified in mammal cancer research). These proteins control the division,
differentiation and migration of cells as well as the specification of the main body axes.
In this thesis, we especially consider the head-tail polarity in flatworms associated with
Wnts.
1.1.3. From signals to genes and back
The function of a cell is largely determined by the proteins inside (244). There are
several classes of proteins. Housekeeping proteins for the maintenance of the basic
cellular functions such as protein synthesis, structural support and cell metabolism are
present in all cells under physiological conditions. Other proteins are only found in
certain types of cells and are involved in specific tasks such as division, developmental
signaling, force generation, sensory perception or immune responses.
The blueprint for all proteins is chemically encoded in the genome. The genome
consists of Deoxyribonucleic acid (DNA) macromolecules, in which characteristic base
pair sequences (genes) represent the proteins. This genetic information can be read out
from the DNA strands by a gene expression pathway as depicted in Fig. 1.4(a). First,
the DNA unfolds at the respective gene site and the gene sequence is successively copied
6
(a)(b)00.250.50.751-1001020304050RelativefluorescenceDistance to source boundary [μm]Stage 1 (0 h)Stage 3 (1.5 h)Stage 4 (2 h, midplane)Stage 4 (2 h, surface)100 μm1.1 Development, growth and regeneration
Figure 1.4.: (a) Gene expression: proteins are synthesized by transcribing the genetic
information saved in the DNA to mRNA molecules, which then act as a blueprint
for the protein.
(b) Gene expression can be activated or de-activated by signaling
molecules (here exemplified for canonical Wnt signaling).
In the absence of Wnt
molecules, β-catenin is tagged for degradation by a destruction complex, which in-
cludes Axin, Adenomatosis polyposis coli (APC), Glycogen synthase kinase 3 (GSK3)
and Casein kinase 1α (CK1α).
If Wnt is bound to the Frizzled receptors and the
Low density lipoprotein receptor-related protein 5 or 6 (LRP5/6), the formation of
the destruction complex is suppressed and β-catenin can translocate to the nucleus to
act as a co-activator of various genes (68, 143, 244).
to a messenger ribonucleic acid strand (mRNA) during a process called transcription.
Again a sequence of nitrogenous bases encodes the specific protein. Second, during
translation, the mRNA acts as a template for protein synthesis with the help of
ribosomes (i.e. large complexes consisting of proteins and RNA strands).
All somatic cells (with a few exceptions) are genetically equivalent because they all
stem from the same initial egg cell. During cell division, the DNA becomes duplicated
and one copy remains in each daughter cell (68, 244). The cells acquire different fates
if different genes are activated and thus different proteins are present in the cells.
This activation of genes is in turn also controlled by signaling molecules. Fig. 1.4(b)
illustrates the signaling cascade of canonical Wnts as an example. In the absence of
Wnt molecules, β-catenin is tagged for degradation by a destruction complex involving
several molecules such as Adenomatous poluposis coli (APC) and Axin. Upon binding
of Wnt molecules to the Frizzled receptors and co-receptors (LRP5/6), Axin is recruited
to the membrane and the formation of the destruction complex is suppressed. Thus,
β-catenin can accumulated and reach the nucleus, where it acts as a transcription
co-activator for specific target genes.
In effect, Wnt has an activating effect on the
expression of these target genes.
The resulting proteins can fulfill certain tasks for the cell in which they have been
7
DNAmRNAnucleustranscriptionproteinribosometranslation(a)(b)target genetarget geneWnttranscriptionrepressorstarget geneWntFrizzledLRP5/6β-cateninβ-cateninβ-cateninβ-cateninβ-cateninDishevelledAxinAxinAPCβ-cateninGSK3CK1αtranscrip.factorstranscrip.factors1. INTRODUCTION
synthesized, yet they can also be released and act as morphogens to activate or de-
activate parts of the genome in other cells. This realizes positive and negative feedback
loops, from which complex cellular signaling networks are built.
1.1.4. Gene expression can be modified in experiments
Experimentalists can interfere with the synthesis of specific proteins by exploiting the
control and error correction machinery of the cells, which ensures robustness of the
important processes of DNA duplication, transcription and translation and modifies
their outcome (68). One such mechanism is RNA interference (RNAi), for which small
interfering RNA (siRNA) pieces are used by the cell to target specific mRNA strands,
mainly for destruction (6). For example, this can be an immune response against exoge-
nous RNA introduced by viruses. Similarly, experimentalists can artificially suppress
a protein of choice by introducing a RNA sequence for this protein in the cell. Such
RNAi techniques are applied to obtain some of the data presented in this thesis.
1.2. Flatworms as a model organism for scaling, growth and
regeneration
Classic experiments on flatworm regeneration already inspired the idea of morpho-
genetic gradients (36, 128, 129). In recent years, the flatworm Schmidtea mediterranea
(Smed ) has become increasingly popular as a model organism to study regeneration and
growth, aging, and even behavior (63, 140, 151, 158, 162, 179, 183, 198, 212). There
are many different flatworm species populating very diverse habitats all around the
world. They are found in saltwater, in freshwater and in the soil; some live more than
1000 m under the sea and some parasitic species (like flukes and tapeworms) in the
body of other organisms (151, 170, 179, 202). Smed is a non-parasitic flatworm living
in freshwater. Such free-living species are sometimes also referred to as "planarians"
(50).
Flatworms (Greek: Platyhelminthes) bridge the gap between other model organisms
of lower complexity such as the freshwater polyp Hydra and the C. elegans worm and
those of higher complexity, such as fruit fly, clawed frog, axolotl, zebrafish and mouse
(184, 244), see Fig. 1.5. Many members of the flatworm phylum seem to represent the
most evolved organisms that are still able to regenerate any part of their body (50, 128,
140, 184). For example, Smed can restore its complete body plan from amputation
8
1.2 Flatworms as a model organism for scaling, growth and regeneration
Figure 1.5.: Flatworms are the most complex model organisms that can still regene-
rate every tissue. This regeneration capability is shared with simpler organisms such as
Hydra. At the same time, flatworms possess organ systems like a centralized nervous
system and two distinct brain lobes, which are characteristic for the most complex
organisms (50, 68, 128, 140, 143, 184, 244).
fragments as tiny as the very tip of the tail with only about 104 cells (126). By
re-growing missing body parts and re-modeling oversized organs, they recover their
normal shape scaled to the size of the amputation fragment within 1-2 weeks (4, 170).
This astonishing regeneration capability is comparable to the much simpler Hydra and
distinguishes flatworms from other model organisms with much less body plan plasticity
(158, 184).
Thereby, the body plan of Smed shows already key characteristics that are usually asso-
ciated with higher order organisms (140), see Fig. 1.6. The bilaterally symmetric Smed
possess a central nervous system with a distinct bilobed brain and two ventral nerve
cords connected by commissural neurons (69, 95, 165, 170, 171, 183). The sensory sys-
tem processes information from chemo-, rheo- and photorecetors leading to a complex
behavioral repertoire (69, 95, 153, 171). Their usual mode of motility is a gliding motion
on a secreted layer of mucus being propelled by the beating of numerous short flagella
(or cilia) that project from their multi-ciliated ventral epithelium (8, 169, 179, 180).
However the worms also use the numerous muscles situated along their inner body wall
for (i) steering, (ii) quick escape responses, (ii) exploratory head motion and (iv) a
back-up movement strategy (152, 153, 169, 180, 220).
9
AxolotlZebrafishClawed frogFruit flyC. elegansHydraMouseFlatworm"Complexity"Regeneration capabilityHydra:regenerates even from cell aggregatesno regeneration, rigid lineagevery limited regeneration in the embryoregeneration very limited and only in tadpolesregeneration of major body parts like limbs, heart, spinal cord, brainregeneration of major body parts like fin, heart, spinal cordregeneration limited to some tissues like liverregeneration of all body partsC. elegans:Flatworm:Fruit fly:Clawed frog:Axolotl:Zebrafish:Mouse:1. INTRODUCTION
Figure 1.6.: The body plan of the flatworm Schmidtea mediterranea (Smed ) shows
already key characteristics that are usually associated with higher order organisms
(37, 51, 171, 173, 231, 245).
Smed belong to the taxon Tricladida, which is reflected by the fact that their highly
ramified gut splits in three main branches (69, 140, 165, 170), see Fig. 1.6(b). During
feeding, the carnivorous worms suck in food through their extensible pharynx opening.
After digestion, the pharynx also functions as an anus for excretion. Protonephridia
constitute a further part of the excretory system, which performs similar functions as
the kidneys in humans (171).
Turnover, growth and regneration completely relies on a pool of stem cells called
neoblasts. The fraction of neoblasts among all cells had been estimated to be as large
as 25− 35% (15, 16, 17, 18). If the worms are depleted of all neoblasts by γ-irradiation,
they will show a regression of the body starting from the head and finally they fall
apart (15, 46, 159, 170). The time scale until the decay sets in varies between species
in the range from several days to a few weeks (46, 159). This indicates that there is no
de-differentiation of cells to restore the stem cell pool. A subpopulation of neoblasts
is pluripotent (and maybe even totipotent) and can develop into every other cell type
(15, 63, 84, 140, 158, 170). Wagner et al. have shown that irradiated worms can be res-
cued by transplanting a single pluripotent neoblast from an intact worm (229). Even
though neoblasts are well defined by their progression through the cell cycle resul-
ting in cell division, there is increasing evidence that the neoblast population is not
10
(a) locomotory and sensory system(b) digestive and excretory system(c) hermaphrodite reproductive systemeye spotsbrain lobesventral nerve cordscilia(mostly ventral)pharynx (ventral)gutprotonephridiaovariestestescopulatory organs1.2 Flatworms as a model organism for scaling, growth and regeneration
homogeneous (85, 132, 194, 225). Most likely, they also comprise lineage-restricted
subpopulations and transiently amplifying cells that go through a few more rounds of
cell division during differentiation.
Smed show both sexual and asexual reproduction (48, 140, 170, 183). Some strains are
hermaphrodites, which develop testes and ovaries symmetrically along their body axis
as well as copulatory organs for cross-fertilization, see Fig. 1.6(c). In contrast, asexual
strains do not possess a reproductive system and
reproduce by fissioning: the worm attaches its tail
to the substrate and glides on until the body is
ripped into two or more pieces, which develop into
new worms after regeneration (140). Fissioning
depends strongly on the environmental conditions,
in particular on temperature, feeding, light and
worm density (48, 131, 140, 161). For example,
fissioning frequency is reduced in crowded envi-
"Wenn einem beim Duell ein Ohr oder sonst
ein Glied abgeschlagen wurde, so wuchs inner-
halb von acht Tagen erstens ein neues Ohr an
den Menschen und zweitens ein neuer Mensch
an das Ohr.
... Wer sich vermehren wollte,
schnitt sich zum Beispiel einen oder zwei oder
zehn Finger ab und wartete acht Tage lang."
-- Joachim Ringelnatz, Abseits der Geogra-
phie, 1924 (168)
ronments. Like regeneration abilities, reproduction strategies also vary widely among
flatworm species (50). In particular, one observes various approaches to asexual repro-
duction. Some species like Smed first split and regenerate afterwards, others already
grow the respective organs of the new body plan before fissioning. The second stra-
tegy comes in two forms: as paratomy with the new body aligned to the old axis and
budding with a non-aligned outgrowth at the side or pointing backward.
Smed are well-suited model organisms to study growth and cell turnover. First, all
somatic cells are continuously replaced by cell turnover at a time scale of weeks (158,
170). Second, growth and cell turnover rely completely on the division of a large
population of neoblasts, some of which are pluripotent (15, 63, 85, 140, 158, 170). Third,
depending on feeding conditions, the worms can grow and actively shrink (usually
referred to as "degrow"), while scaling their body plan over more than one order of
magnitude in length in the range of 0.5 mm to 2 cm (140, 151, 170), see Fig. 1.7(a).
Active degrowth enables the worms to survive long starvation periods of several months.
It has been suggested that worms recycle apoptotic material during degrowth to fuel
the metabolism of the remaining cells (69, 70).
Smed are also well-suited to study scalable patterning during growth and regeneration.
Being able to reversibly grow by a factor of 40 in length and perfectly regenerate
even from scrambled body fragments prompts for patterning mechanisms that are not
only highly robust and self-organizing but also functional across a wide range in sizes,
11
1. INTRODUCTION
Figure 1.7.: Growth and regeneration in flatworms:
(a) Schmidtea mediterranea
(Smed ) can reversibly grow over a 40fold range depending on feeding conditions (Im-
ages taken by Nicole Alt under the supervision of the author). (b) Amputation frag-
ments regenerate to form a perfectly shaped worm within 2 weeks (green lines mark
the cuts, white tissue parts indicate the regeneration site before all pigments have been
reestablished, modified with permission from (116)).
see Fig. 1.7. For example, regeneration comprises a tightly controlled sequence of
responses (4, 10, 13, 14, 69, 81, 117, 128, 158, 165, 170, 237). After an immediate muscle
contraction to close the wound, a peak in cell division together with the migration of
stem cells within the first 12 h generates an outgrowth of undifferentiated tissue at the
wound site, the blastema. In a second step, the now oversized, remaining body parts
are re-patterned to match the size of the amputation fragment. The corresponding
process is called morphollaxis and comprises a more sustained proliferation and cell
death response during the following days. All this pattern formation is guided by
an internal coordinate system that is re-established as one of the earliest cues in the
remaining worm fragment independently of cell differentiation (81, 141).
Our work is especially focussed on the anterior-posterior (AP) axis specifying head
and tail position. The Wnt/β-catenin system is a conserved pathway for such AP
patterning in many animals and it is also key for the AP polarity in flatworms (4, 9,
54, 81, 82, 141, 160). There are 9 Wnt genes in flatworms. In particular the canonical
Wnts (activating β-catenin signaling) are highly expressed in the tail.
Inhibition of
the Wnt/β-catenin system leads to formation of additional heads, while overexpression
induces tail identity everywhere in the worm (3, 82, 82, 160). Remarkably, reducing
the level of β-catenin is sufficient to induce full regenerative power to some flatworm
species that are usually deficient in head regeneration (116, 200, 222). Taken together,
the experiments suggest that β-catenin is instructive for positional information along
the AP axis in a concentration-dependent manner.
12
1mm(a)(b)Growth and DegrowthRegeneration1.3 Theories of body plan patterning by morphogens
Perpendicular to the AP axis, the dorsal-ventral (DV) axis and the medial-lateral (ML)
axis are formed (54, 117). For example, the DV axis is specified by the interplay between
the TGF-β family members Bmp and Admp. Together, all three body axes are charac-
terized by gradients in abundance and expression of specific molecules, in neoblast
mitotic activity as well as in membrane potential (4, 9, 23, 62, 81, 149, 201) and it is
a major open question how these gradients robustly guide body plan patterning irre-
spective of the size of the organism.
1.3. Theories of body plan patterning by morphogens
Here, we will discuss two ideas of pattern formation based on cell-cell communication
via morphogens. The first type of mechanisms describes how a pre-existing structure
can act as an organizing center for the development of further patterns. Such theories
are built on the idea of a well-defined, sequential developmental program where existing
patterns determine the formation of new patterns, see Fig. 1.8(a). Maternal patterning
cues provide a pre-pattern for the embryonic tissues forming from an egg cell and the
layout in the embryo determines the future body plan of the animal (2, 57, 172). For
example the localization of bicoid mRNA to the anterior pole of the fly embryo is a
maternal effect. However, the impressive capabilities of flatworms to regenerate from
almost arbitrary initial fragments challenge these concepts, see Fig. 1.8(b). In contrast
to the sequential patterning from predefined cues, Alan Turing introduced a second
class of mechanisms for self-organized pattern formation (221). We will discuss both
approaches and assess them from a biological perspective.
Body plan patterning requires that cell fates are assigned depending on the relative
position of the cells in the tissue or the organism (68, 244).
In order to obtain the
information about their spatial position, cells sense their environment and communicate
with each other. For that, cells respond to mechanical cues as well as chemical signals
such as morphogens (89, 112, 147, 174, 224).
Signaling molecules often establish long-range, graded concentration profiles, which can
be accounted for by the interplay of transport and degradation (76, 223, 232). Cells
respond to the concentration of these molecules in their local environment. Thus, a
graded morphogen concentration can provide cells with the information about its spatial
distance from the morphogen source (243), see Fig. 1.9(a). This idea forms the basis for
the French flag model, which draws a simplified picture of how body plan patterning
might be guided by graded morphogen profiles. For example, a stripe pattern can be
13
1. INTRODUCTION
Figure 1.8.: (a) Idea of sequential development: The emerging tissues in the fertilized
egg determine the future body plan of the juvenile chicken and later of the adult
hen. Finally maternal signals break again the spatial symmetry in the new egg and
guide development of another chicken. (b) Regeneration in flatworms: Diverse initial
fragments are repatterned to form the same body plan scaled to fragment size.
generated if cellular differentiation depends on discrete, genetically encoded threshold
levels. In Fig. 1.9(a), cells turn blue if the morphogen concentration is above the first
threshold level and red if it is below the second threshold level. Several intersecting
gradients can generate more complex patterns.
Dose-dependent responses have indeed been observed in experiments, yet in a more
complex way than given by the simplified French flag model (72, 74, 75, 80, 104, 191,
242). For example, cells of the frog Xenopus show a distinct differentiation response to
at least five concentration thresholds of Activin (72, 75). However, the early response
of the individual cells is less specific and very inhomogeneous (73, 242).
It appears
that only the interaction between cells leads to the well-defined behavior of the cell
aggregate. Furthermore, the cells in the tissue are reported to respond in a ratchet-like
manner to the highest level of Activin to which they were exposed (80). Similar dose-
dependent responses have also been recently reported for Wnts in the frog embryo with
respect to the AP axis patterning (104).
Thus, although the simple concept of a direct threshold-comparison has been questioned
in some organisms, the French flag model continues to provide a pictorial representa-
tion of the case where a feedback on the concentration gradient by the responding cells
can be neglected. As an extension, it also has been proposed that the cells might com-
pare spatial and temporal differences in morphogen concentration and correspondingly
adjust cell division, differentiation and motility (135, 167, 175, 233, 234, 235).
The described mechanisms can explain patterning, given a pre-defined source region
where morphogens are produced locally. Yet, how is this source region established
in the first place? One possibility is that another graded morphogen profile specified
14
(a)(b)1.3 Theories of body plan patterning by morphogens
Figure 1.9.: (a) The French flag model substantiates the idea that a graded mor-
phogen profile provides positional information for cells in a tissue. Specifically, cells
adopt distinct cell fates (sketched blue, white, red) by responding to the local mor-
phogen concentration depending on whether a certain threshold is met. The model
assumes a pre-defined source region (grey), which secretes the morphogens. (b) In
a Turing system, at least two chemical species (morphogen 1 and 2, red and blue)
interact. Source regions (gray) of the morphogens are established in a self-organized
way (here: a source forms where the concentration of the first morphogen is larger
than the concentration of the second morphogen).
this source by a threshold rule and a sequential developmental program establishes one
pattern from an already existing one, see Fig. 1.8(a). An alternative explanation dates
back to Alan Turing. In a seminal paper from 1952, he proposed a general framework
for the spontaneous formation of biological patterns, independent of pre-patterning
cues (221). Turing's framework was later further explored by Meinhardt and Gierer
(64, 65, 66, 106, 120).
Turing demonstrated how the interaction of at least two diffusing molecular species
can result in chemical patterns in a completely self-organized manner, see Fig. 1.9(b).
Thereby, these (often periodic) patterns specify their own production regions. It is an
appealing idea that the chemical patterns layout the body plan of an animal. Yet for
the following 50 years, there was only little experimental evidence for Turing's ideas
and the focus of developmental biology was shifted to other patterning mechanisms
like the French flag model (74). Certainly this was partly due to the fact that it is
generally difficult to demonstrate experimentally that a specific pattern is generated by
a Turing mechanism. While a unidirectional relationship like in the French flag model
(where a concentration of one molecule has a particular effect on other molecules or the
cells) is straight forward to analyze, this is less so for Turing patterning which includes
cross-reaction terms. The behaviour of a Turing system is often counterintuitive and its
analysis might easily yield misleading results. For example, the concentration of some
15
morphogen concentrationthreshold level 1threshold level 2(a)(b)Pre-patterned: morphogencellsSelf-organized:morphogen 1morphogen concentrationsmorphogen 21. INTRODUCTION
Turing molecules peak at the maximum concentration of their repressors, compare to
Section 1.4. Furthermore, if there are more than two chemical species involved, the con-
centration of one of them might decrease upon knockout of its direct activator because
of additional indirect effects, see Appendix A.7.4 for an example. Despite the chal-
lenging task of revealing the existence of Turing patterns and identifying the involved
molecules, recently more and more evidence has been accumulated that Turing's ideas
might be the guiding principles for the formation of a wide range of biological patterns,
ranging from the formation of digits in vertebrate limbs to the emergence of left-right
asymmetry (49, 119, 134, 137, 164, 191, 197). Some of these examples combine Turing
patterning with a French flag model.
In fact, in many biological systems, pattern formation might result from a combination
of both concepts (74). The feedback loop of a Turing mechanism ensures high robust-
ness and possesses the ability to generate chemical patterns from random fluctuations.
In particular, Turing systems can spontaneously generate graded concentration profiles
like required for the French flag model. In turn, biological patterns hardly emerge in
a completely homogeneous environment without any pre-patterning cues, as Turing al-
ready remarked (221). During development as well as during regeneration, pre-existing
morphogen profiles and tissue structures can guide the formation of chemical patterns.
Thus, Turing mechanisms might often be found downstream of morphogen profiles that
modify the respective patterns (74, 164).
1.4. Turing mechanism yields self-organized patterns
In 1952, Alan Turing introduced a generic framework for the self-organized formation
of chemical patterns in biology (221). He asked the very general question whether there
are conditions, under which a system of two or more diffusing and reacting chemical
species possesses a homogeneous steady state, which is linearly unstable with respect to
inhomogeneous perturbations, such that inhomogeneous patterns form spontaneously.
Very importantly, in consequence, any model for self-organized patterning based on
diffusing and cross-reacting molecules can be understood within the Turing framework
(64, 65, 120, 147, 195, 221). Recently, there are attempts to further generalize the
Turing mechanism to generic interactions between two players without a diffusion term
(136, 236).
In the following we show that the conditions for a Turing instability can be derived
from the linear stability analysis of the homogeneous steady state. For this, we briefly
16
1.4 Turing mechanism yields self-organized patterns
recall the basic Turing model comprising two chemical species with concentrations A
and B. Details on the derivations are provided in Appendix A.3 or in the literature
(64, 65, 146, 195, 221). The general reaction-diffusion system for two molecular species
in one space dimension is
∂tA = RA(A, B) + DA ∂2
∂tB = RB(A, B) + DB ∂2
x A
x B ,
(1.1)
with diffusion coefficients DA and DB and two generic functions RA and RB describing
the reactions between the different molecules and the effects of each molecular species
on itself.
In order to spontaneously form stable patterns from random fluctuations,
h, B∗
these reaction functions have to fulfill two requirements: the homogenous steady state
(A∗
h) = 0, has to be (i) stable with
respect to homogeneous perturbations (to avoid a diverging behavior), yet (ii) unstable
h), defined by RA(A∗
h) = 0 and RB(A∗
h, B∗
h, B∗
with respect to inhomogeneous perturbations.
It is possible to derive a set of necessary conditions for spontaneous pattern formation
from these two conditions above by applying linear stability analysis. For this, we
consider a small perturbation (a, b) about the homogeneous steady state: A = A∗
and B = B∗
h + b. The linearized form of Eq. 1.1 can be written as
h + a
.
(1.2)
(cid:32)as
(cid:33)
∂t
bs
= Ms
(cid:32)as
(cid:33)
bs
(cid:90)
(cid:90)
Here, the perturbation modes as(t) and bs(t) with wavenumber s represent the spatial
Fourier transform of a(t, x) and b(t, x), respectively:
as(t) =
a(x, t) e−2πsx/L dx ,
bs(t) =
b(x, t) e−2πsx/L dx
(1.3)
The matrix Ms is given by
(cid:32)∂ARA − DA(2πs/L)2
∂ARB
(cid:33)
,
(1.4)
∂BRA
∂BRB − DB(2πs/L)2
Ms =
h, B = B∗
h.
where derivatives are evaluated at A = A∗
In order to fulfill the two conditions on the stability above, (i) the real parts of both
eigenvalues qI
s of the matrix Ms have to be negative for s = 0 and (ii) at least
one has to be positive for s (cid:54)= 0. As shown in Appendix A.3, this results in the following
conditions for the trace Tr and the determinant Det of the matrix Ms:
s and qII
Tr[M0] < 0 , Det[M0] > 0
and ∃ s > 0 with Det[Ms] < 0 .
(1.5)
17
1. INTRODUCTION
Figure 1.10.: There are two possible feedback topologies that lead to pattern for-
mation in a Turing system with two molecular species (A and B): (a) an activator-
inhibitor scheme with activator A and inhibitor B, (b) a second topology where both
chemical species have activating and inhibiting effects. We show two examples of
typical concentration profiles corresponding to the two feedback topologies. The self-
organized source regions are indicated in gray.
In consequence, we obtain several constraints on the reaction design, which have been
summarized by the principle of local activation and lateral inhibition (64, 65, 120, 146,
147), see Appendix A.3 for details and Fig. 1.10 for illustration of the allowed feedback
topologies. The necessary conditions for spontaneous pattern formation are:
1. One molecule has to be self-activating and the other self-inhibiting. In the following,
we choose A to be the self-activator and B the self-inhibitor:
2. There have to be cross-reaction terms of opposing sign:
(cid:16)
∂ARB
∂BRA
< 0
∂ARA > 0
and ∂BRB < 0 .
(cid:17)(cid:16)
(cid:17)
(1.6)
(1.7)
3. The diffusion coefficient of the self-activator has to be sufficiently smaller than the
diffusion coefficient of the self-inhibitor:
DA < DB .
(1.8)
As a result for a Turing system with two chemical species, there are two possible network
topologies as depicted in Fig. 1.10. In Turing feedback 1, the concentration A has only
18
(a)(b)Morphogen concentrationPositionPositionsource of A & BMorphogen concentrationsource of AFeedback 1Feedback 21.5 Open questions in growth control and scalable body patterning
activating effects and B has only inhibiting effects both on itself and the respective
other player. In this case, we can shortly refer to A and B as the concentrations of
activator and inhibitor instead of self-activator and self-inhibitor. In Turing feedback 2,
both molecular species have activating and inhibiting effects. This includes depletion
models, where, for example, binding of both molecular species enhances the production
of A but consumes B (66). Note that there is a formal correspondence between the two
Turing topologies by the replacement B → Bc − B with a constant parameter Bc.
How the principle of local activation and lateral inhibition leads to spontaneous pat-
tern formation can be nicely demonstrated for the Turing feedback 1 of Fig. 1.10(a).
At the source, the activator dominates because the inhibitor is diffusing away more
quickly. Consequently, the source region stabilizes itself and initially tends to expand.
In contrast, some distance away from the source, the inhibitor dominates due to its fast
diffusion. This creates at least two distinct regions, which already comprises a simple
pattern. In contrast to other systems where diffusion homogenizes a pattern, here dif-
fusive spreading in combination with specific reactions enhances small inhomogeneities
in the concentrations. Therefore, it is sometimes also referred to as a diffusion-driven
instability (145, 146, 181).
1.5. Open questions in the study of growth control and
scalable body patterning
During development, growth and regeneration, cells communicate with each other and
mutually influence their behavior, especially by changing the expression status of their
genes. This information exchange orchestrates cell division, cell death and differen-
tiation. It can also guide cell migration and can elicit the release of further signals,
e.g. the secretion of morphogens. As a common theme of this thesis, we devise theo-
retical descriptions of how cell fate decisions based on local rules on the microscopic
scale result in the formation and maintenance of a macroscopic body plan, drawing
inspiration from flatworms as a specific model organism. In particular, we address the
following questions:
How is the turnover of cells regulated during growth and homeostasis and
how are fluxes in cell number related to fluxes of metabolic energy? Organisms
have to control the number of cells to ensure homeostasis and growth in a well-defined
manner. As changes in cell number depend on the balance between cell proliferation
and cell loss, there needs to be a communication between dividing and differentiated
19
1. INTRODUCTION
cells (158). Furthermore, the environment and in particular the availability of food
provides an external stimulus to influence cell behavior, such as proliferation and cell
death (113, 244).
What are minimal requirements for self-organized patterns that scale with
organism size? Organisms also have to control the type and position of cells for a
reliable body plan patterning. Again, this requires a communication between cells,
ranging from direct neighbor-neighbor interactions to long range signaling via mobile
molecules such as morphogens (54, 68, 244). Importantly, pattern formation can be
observed on all length scales from the development of a fertilized egg to growth and
regeneration of large scale tissues in mature organisms. Yet, patterning mechanisms of-
ten possess fixed characteristic length scales defined by the intrinsic physical properties
of the system (76, 223, 232), which is challenged by cases of biological pattern scaling
like regeneration in flatworms.
We aim to understand scalable body plan patterning and cell turnover dynamics by
building on the framework of dynamical systems theory and birth-death processes.
Biological systems add a new perspective to those classical concepts. They usually
operate far from equilibrium and form patterns with unconventional properties (39,
87, 120). Development and regeneration are subject to a high level of noise ranging
from external perturbation like a variable environment to the intrinsically stochastic
nature of gene expression (77, 218, 219). Thus, robust growth control and body plan
patterning require reliable sensing, transmitting and processing of noisy information
(47, 218). We address the question of robustness with respect to initial conditions and
physical parameters as well as the structural robustness of the models.
This thesis is mainly inspired by the astonishing scaling and regeneration capabilities
of flatworms, yet the results obtained here are likely also to be relevant for other bio-
logical organisms as well as related questions in biological physics, nonlinear dynamics
and pattern formation.
20
1.6 Organization of the thesis
1.6. Organization of the thesis
In this thesis, we investigate biological shape and size control on various levels, for which
we combine theoretical descriptions and state-of-the-art analyses of experimental data
in flatworms.
First, we explore mechanisms for self-organized and self-scaling pattern formation
(Chapter 2). We discuss to what extent previously proposed theories can account
for the scaling of morphogen profiles. For this, we adhere to a strict mathematical
definition of scaling and illustrate the difference between perfect scaling and approxi-
mate scaling with system size. Furthermore, we demonstrate the absence of scaling in
classical Turing pattering, going beyond linear stability analysis.
Second, flatworms like Smed challenge existing theories on body plan patterning, which
typically can at most only explain one of both: either scaling or self-organization. Here,
we introduce and characterize a novel class of mechanisms that combine both features
(Chapter 3). The developed theory can act as a framework to understand robust body
plan scaling during growth and regeneration in flatworms.
Next, worm shape variations cannot only be observed between different stages of de-
velopment, growth and regeneration. Individual worms also show large body deforma-
tion during movement, driven by muscle contractions. We apply a method based on
"Principal component analysis" to characterize shape dynamics and analyze movement
patterns. A similar shape mode analysis also enables us to compare shapes of different
related species. Analyzing motility patterns and comparing shape variations between
worm species is the first step towards relating form and function (Chapter 4).
Finally, we analyze growth and degrowth dynamics and their dependence on feeding
conditions. We investigate mechanisms for size control by metabolic energy balances
to explain the macroscopic growth and degrowth behavior in terms of microscopic cell
turnover dynamics. Our theory makes testable predictions for the ongoing experiments.
Beyond, we develop the theoretical basis for additional measurements on various scales
to reveal further details about the control logic of cell turnover (Chapter 5).
21
2. Scaling in morphogen systems
2.1. Scaling of biological patterns
The development of multi-cellular organisms with a well-defined, complex body plan is
one of the most fascinating processes in nature. A series of patterning events take place
"The continuous change in form that takes place
from hour to hour puzzles us by its very simpli-
city. The geometric patterns that present them-
selves at every turn invite mathematical analy-
sis." -- Experimental embryology, Thomas H.
Morgan, 1927 (130)
across various length scales leading to a distinct
layout, which is scaled to match the size of the
organism. The scaling of body plan patterns be-
comes especially apparent if a juvenile organism
already resembles its adult counterpart or if indi-
viduals of different, yet related species look very
much alike besides their great differences in size (27, 76, 223). A third, more subtle
example is the robust formation of proportionate patterns in the same organism during
development despite size variations that arise from varying environmental conditions
or stochastic fluctuations of growth rates (24, 76).
In this chapter, we analyze to what extent previously proposed theories can account
for pattern scaling. First, we introduce a mathematical definition of gradient scaling.
Based on these considerations, we revisit and assess scaling mechanisms proposed for
morphogen gradients in pre-patterned systems and extract the main principles under-
lying scaling. This will later allow us to point out the important differences between
scaling mechanisms for pre-patterned and self-organized systems in Chapter 3. Fur-
thermore, we discuss the absence of scaling in self-organized Turing systems beyond
the classical approach of linear stability analysis. The latter has been published in
Werner et al. (239).
22
2.2 Morphogen dynamics and concentration profiles
2.2. Morphogen dynamics and concentration profiles
One of the most simple morphogen systems that result in graded concentration profiles
draws on diffusion from a localized source with an effective diffusion coefficient D and
a linear degradation with rate β (232). It has been frequently applied to describe mor-
phogen gradients in the fruit fly (24, 44, 76, 78, 234). Note that effective diffusion might
result from a wide range of underlying undirected processes, such as active transport
or even signaling between neighboring cells without secretion of motile molecules.
For simplicity, we consider a one-dimensional system of size L with reflecting boundary
conditions. The corresponding time evolution of the morphogen concentration C =
C(t, x) is given by
x C − β C + ν
∂tC = D ∂2
∂xCx=0 = ∂xCx=L = 0 .
(2.1)
(2.2)
Here, ν(x) = α Θ(w − x) describes localized morphogen production with rate α in a
source of width w. Θ denotes the Heaviside step function.
Eq. 2.1 also holds in two and three dimensions, if the system is symmetric with respect
to the other dimensions and, thus, the morphogen concentration still only depends on
x, see Appendix A.1.1. If the tissue grows at a time scale comparable to the time scale
of the morphogen dynamics, one has to consider additional terms for advection and
dilution.
The steady state solution to Eq. 2.1-2.2 is computed in Appendix A.1.2. Here, we are
mainly concerned with the steady state concentration outside the source region, which
is given by
with amplitude
C∗(x) = C0
cosh(L/λ − x/λ)
cosh(L/λ)
for w ≤ x ≤ L
and a characteristic length scale
C0 =
α sinh(w/λ)
β tanh(L/λ)
λ =(cid:112)D/β .
(2.3)
(2.4)
(2.5)
Within the course of a characteristic time scale (which is of the order of 1/β in our
example, see Appendix A.1.3), the morphogen will relax towards this steady state pro-
file. Analyzing the steady state is well justified in a case of separated time scales,
where other dynamics (e.g. growth and differentiation) are slow in comparison to the
relaxation time of the morphogen profile. In the following, we will often assume such
23
2. SCALING IN MORPHOGEN SYSTEMS
a quasi steady state. One should keep in mind that this is a simplifying assumption
which is not always fulfilled. In fact, it has been argued that many gradients might be
already read out before the steady state is reached as a means to increase robustness
(20, 29, 55, 167). Yet, even in these cases, the steady state is the reference state and
provides a first order approximation to the concentration the cells actually respond to.
2.3. Scaling of concentration profiles with system size
Profiles of signaling molecules have been measured for example in the fruit fly Droso-
phila melanogaster, which is an important model organism to study biological pattern
formation (123, 206). Quantifications of morphogen concentrations in the developing
wing and eye of the fly at different stages of development revealed that the concentration
profiles maintain an approximately constant shape relative to the changing size of the
growing tissue (26, 83, 234, 235). This biological observation of profiles that scale with
system size inspired the development of a number of theoretical mechanisms that can
account for pattern scaling (11, 24, 26, 27, 96, 135, 148, 154, 223, 233, 234). In this
section, we discuss examples, which represent two major classes of these mechanisms.
We define a mathematical property of perfect shape scaling and analyze approximations
of this property.
We define perfect shape scaling as the ideal case of strictly proportional scaling of a
concentration profile with system size. In this case, the concentration at steady state
can be written as
C(x; L) = C0(L) · Z(x/L) ,
(2.6)
where C0(L) denotes a possibly size-dependent amplitude and Z(x/L) defines the shape
of the profile, which only depends on the relative spatial coordinate x/L. As a result,
scaling profiles from systems of different sizes perfectly collapse onto a single master
curve if plotted as a function of relative coordinates and normalized by their amplitude,
see Fig. 2.1(a). For the steady state solution in Eq. 2.3, perfect shape scaling arises
if the characteristic gradient range λ is not constant but strictly proportional to the
length of the system L:
λ ∝ L .
(2.7)
In a more general case, if the length scale λ increases with L in a monotonic, yet
nonlinear fashion, we will classify this as approximate scaling, see Fig. 2.1(b):
∂Lλ > 0 .
(2.8)
24
2.3 Scaling of concentration profiles with system size
Figure 2.1.: Illustration of the mathematical concept of perfect shape scaling and
approximate scaling: (a) Perfect shape scaling: normalized concentration profiles col-
lapse onto a master curve when plotted as a function of the relative position. (b) In
a more general case, concentration profiles might expand with system size, though
the collapse onto a single master curve is only approximate. We classify this case as
approximate scaling.
In the following, we will strictly distinguish between the two notions of approximate
scaling and perfect shape scaling. This slightly academic distinction will prove use-
ful in the subsequent mathematical analysis. Nonetheless, we want to emphasize that
approximate scaling can be virtually indistinguishable from the ideal case of perfect
shape scaling, both in measurements and simulations. In contrast to our theoretical
description, real systems are generally prone to intrinsic fluctuations and measurement
inaccuracies. Furthermore, the separation of time scales between system growth and
morphogen dynamics assumed here will always be an approximation, such that the mor-
phogen concentration never exactly corresponds to the theoretically computed steady
state profile. Therefore, even if the underlying mechanism is in principle able to gener-
ate profiles that perfectly scale like in Fig. 2.1, we will never observe a perfect collapse of
the measured concentration profiles. On the contrary, mechanisms of approximate scal-
ing that cannot yield perfect shape scaling in the sense of Eq. 2.6, may still fully account
25
Perfect shape scaling:PositionRel. ConcentrationRel. ConcentrationRelative positionPositionRel. ConcentrationRel. ConcentrationRelative positionApproximate scaling:(a)(b)2. SCALING IN MORPHOGEN SYSTEMS
Figure 2.2.: Size-dependent amplitudes of scaling profiles: Models for the scaling of
morphogen profiles often assume the adjustment of degradation rate β or diffusion
coefficient D with system size L. As a result, the amplitude of the morphogen concen-
tration might change depending on whether the morphogen source size (gray) is fixed
(upper row) or scales with L (lower row). The reference profile (red) is the same in all
plots (Eq. 2.3, α/β = 10, L/λ = 2, L/w = 5). The other profiles correspond to either
half (blue) or twice the system size (green) with constant ratio L/λ and either scaling
or non-scaling source size w.
for gradient scaling as observed in biological systems. As an example for a mechanism
that yields approximate scaling, we discuss a simple realization of a prominent scaling
mechanism, the expansion-repression model. We emphasize that this mechanism has
been successfully applied to analyse profile scaling in fly wing development (24, 26, 27).
According to the definition of λ in Eq, 2.5, scaling can be achieved if either the diffusion
coefficient or the degradation rate or both depend on system size. Two simple choices
are: (i) D ∝ L2, β = const and (ii) β ∝ L−2, D = const.
In Eq. 2.6, we defined scaling of morphogen profiles as a shape property, irrespective of
the amplitude C0(L). This amplitude might be a function of system size L, compare
also to (223). The variation of the amplitude with L can reveal details of the underlying
scaling mechanism, see Fig. 2.2. We discuss this for the two limiting cases of (i) a source
of fixed width w and (ii) a scaling source w ∝ L.
(i) Fixed source: If we increase λ by enhancing diffusion, the amplitude will decrease
because the same amount of morphogen is distributed more homogeneously in the
system, compare to Eq. 2.4 or Eq. A.5.
In contrast, if we increase λ by reducing
degradation, the amplitude will increase due to a longer lifetime of the molecules. If D
26
ConcentrationConcentrationRelative ConcentrationConcentrationConcentrationRelative ConcentrationAbsolute PositionAbsolute PositionRelative PositionAbsolute PositionAbsolute PositionRelative PositionAdjusting DegradationAdjusting Diffusion(i) Fixed Source(ii) Scaling Source2.4 Scaling of morphogen profiles in pre-patterned systems
and β both depend on size, the amplitude can be kept constant.
(ii) Scaling source: Eq. 2.4 shows that for this case the amplitude does not explicitely
depend on D but depends inversely on β. Scaling by changing the diffusion properties
does not affect the amplitude but scaling by adjustment of degradation leads to a strong
size dependence C0 ∝ L2.
Measurements of the signaling protein Dpp in the fruit fly found a decreasing degra-
dation rate with size as well as the characteristic increase in the amplitude (76, 78,
135, 233, 234). In the clawed frog, both cases have been reported, adjustment of diffu-
sion properties for Bmp as well as adjustment of degradation for the Bmp antagonist
Chordin (25, 94).
Next, we discuss two classes of simple models for the scaling of morphogen profiles,
which represent minimal versions of scaling mechanisms proposed in the literature
(11, 24, 26, 27, 96, 135, 148, 154, 223, 233, 234). We will illustrate that models of
the second class rely on a changing amplitude during scaling. Thus, the amplitude of
the morphogen profile encodes the system size.
2.4. Scaling of morphogen profiles in pre-patterned systems
2.4.1. The concept of an expander as a chemical size reporter
In order to achieve scaling of a morphogen profile in a simple set-up as introduced in
Section 2.2, the length scale of the concentration profile has to couple to the size of
the system. Several mechanisms have been proposed for the expansion of a morphogen
profile with system size L that assume an additional molecular species, often called
"expander" (11, 24, 27, 96, 135, 148, 154, 233, 234).
If the concentration E of such
an expander is a function of L, the interaction of these expander molecules with the
morphogen can result in profile expansion with system size. Depending on whether the
expander concentration increases or decreases with system size, the expander might
enhance or reduce either diffusion or degradation of the morphogen, respectively.
Yet, how does the expander obtain its size-dependent concentration and become a
chemical size reporter? In the following, we discuss two main classes of expander me-
chanisms that have been proposed for the scaling of morphogen profiles. The first class
is characterized by the fact that the expander couples to the system size independently
of the morphogen. The second class subsumes mechanisms that allow for self-scaling of
the morphogen profile by a feedback loop between the expander and the morphogen.
27
2. SCALING IN MORPHOGEN SYSTEMS
We illustrate the scope and the limits of the mechanisms at hand. This allows us to
systematically extract the main principles that lead to robust scaling as applied later
to self-organized systems.
2.4.2. A simple mechanism of gradient scaling:
The expander-dilution model
The most simple idea for how an expander concentration can report on system size is
given by the expander-dilution mechanism (135, 233, 234), see Fig. 2.3(a). It requires a
constant amount of a long-lived expander, which is neither produced nor degraded. As
the system size increases, the expander gets diluted and, thus, the expander concentra-
tion depends on system size. Hence, E ∝ 1/L in a one-dimensional system. However,
such a mechanism is highly vulnerable to the loss of expander and in particular could
not easily cope with amputations of parts of the system.
A similar mechanism can overcome this issue by reading out system size from geo-
metrical features like area-volume ratios. As a specific example, we assume that the
expander of concentration E is degraded everywhere in the one-dimensional system and
produced in a source of constant width wE at the boundary:
∂tE(t, x) = DE ∂2
x E(t, x) − βE E(t, x) + αE Θ(wE − L + x) ,
(2.9)
Again, DE is an effective diffusion coefficient, βE the degradation rate and αE the pro-
duction rate of the expander. In the limit of fast expander diffusion (λE =(cid:112)DE/βE (cid:29)
L), the steady state concentration is given by
E∗ =
αE wE
βE
1
L
.
(2.10)
Thus, the expander level E∗ encodes the system size L, see Fig. 2.3(b). In order for the
morphogen to scale with L, the morphogen degradation could for example be coupled
to the expander as
β ∝ E2 .
(2.11)
The size read-out can also be established differently from the specific choice of Eq. 2.9,
see Appendix A.2.1 for a general derivation. For example, the expander could be
produced everywhere in the system and only be degraded at the boundary, as discussed
by other authors (78, 90, 96, 148, 154).
The mechanism is robust against loss of expander, but it requires λE (cid:29) L irrespective
of system size. Note that a small degradation rate results in a large λE, yet also in a
slow relaxation to the steady state, see Appendix A.1.3. Furthermore, the mechanism
28
2.4 Scaling of morphogen profiles in pre-patterned systems
Figure 2.3.: Schematic illustration of how the concentration of an expander becomes
size-depend. (a) Expander-dilution model: a conserved amount of expander gets di-
luted in a larger system. (b) Dynamic size-sensing in a system with constant influx
and size-dependent outflow of the expander.
relies on a tightly controlled size of the source and the degradation zone. In the example
above, the expander source has to keep its width although the system size changes. This
shifts the problem from scaling the morphogen profile to establishing a source pattern
for the expander. Finally, regeneration of arbitrarily shaped fragments seems unlikely
to agree with the notion of a fixed geometry of source and degradation zone. Therefore,
we next discuss the second class of scaling mechanisms, in which the morphogen itself
controls the expander expression leading to a self-regulating feedback loop.
2.4.3. Perfect shape scaling vs. approximate scaling:
The expansion-repression model
The discussion on morphogen scaling has been enriched by the proposal of a class
of models that adjust gradient shape to system size in a self-organized manner (24,
26). These mechanisms build on a feedback loop between morphogen and expander.
Interestingly, this relates to the engineering framework of an integral feedback control,
where the gradient is adjusted by the expander as long as it does not meet its target
range (19, 24). One prominent example is the expansion-repression model. In order to
illustrate the main principles, we discuss one specific simple realization of this model, in
which the expander production is only turned on where the morphogen level falls below
a threshold Cth like considered in (135), see Fig. 2.4(a). Alternative feedback topologies
are analyzed in Appendix A.2.3 and lead qualitatively to the same conclusions.
29
(a) Expander-dilution with long-lived expander(b) Dynamic size-sensing with short-lived expanderconserved amountconst. influxsize-dep. outflow2. SCALING IN MORPHOGEN SYSTEMS
For our choice, Eq. 2.9 of the expander dynamics is modified to
x E(t, x) − βE E(t, x) + αE Θ(cid:0)Cth − C(x)(cid:1) .
∂tE(t, x) = DE ∂2
(2.12)
In order to close the feedback loop and adjust the morphogen length scales, the expander
has to affect either diffusion or degradation of the morphogen, see Section 2.3.
A simple feedback scheme that yields approximate scaling, though not perfect shape scal-
ing. -- Let us start with a simple realization of the expansion-repression mechanism,
which has been first proposed for gradient scaling in the developing fly wing (24). In
this realization, it is considered that the expander suppresses the degradation of the
morphogen (24, 26, 135). Thus, β in Eq. 2.1 is not constant anymore but obeys
β =
β0
1 + E/Eth
(2.13)
with two constant parameters β0 and Eth. In consequence, increasing expander levels
enlarge the range of the morphogen. Yet, in turn, this suppresses the expander pro-
duction. Eventually, a steady state is reached, in which the expander production is
just cancelled by the total expander degradation in the system. The source size of the
expander in this steady state is determined by
C∗(L − w∗
E) = Cth ,
(2.14)
where C∗ is given by Eq. 2.3 and w∗
E denotes the size of the self-organized source region
of the expander, see Fig. 2.4(a). In the limit of a quickly spreading expander (λE (cid:29) L),
its steady state concentration is
E∗ =
αE w∗
E
βE L
.
(2.15)
From this, we can note the following observations: First, the expander source (estab-
E (cid:54)∝ L), because this
lished by the feedback with the morphogen) must not scale (w∗
would result in a constant expander level irrespective of system size. Second, the mor-
phogen amplitude must change for perfect scaling. Otherwise, if the amplitude were
constant with a scaling gradient range λ ∝ L, we would obtain w∗
E ∝ L from Eq. 2.14
and the expander would not encode the system size. This generally arises from the
fact that in such a mechanism for scaling via a feedback loop both players, morphogen
and expander, have to encode the system size in the steady state. Third, the particu-
lar feedback scheme including Eq. 2.13 yields approximate scaling of the concentration
profiles, though not a perfectly linear relationship of λ ∝ L as required for perfect shape
scaling, see Fig. 2.4(b)-(c).
30
2.4 Scaling of morphogen profiles in pre-patterned systems
Figure 2.4.: A simple realization of the expansion-repression models that yields ap-
proximate scaling, yet no perfect shape scaling. (a) In this specific model the mor-
phogen confines the expander source to the region where the morphogen concentra-
tion is below Cth. (b) Rescaled steady state profiles do not perfectly collapse. (c)
The gradient range λ increases with system size in a sub-linear fashion. We deter-
mine λ by (i) computing λ =(cid:112)D/β at various positions in the system (blue dots),
(ii) fitting the solution of Eq. 2.3 (solid red line) and (iii) calculating it directly in
the limit of λE (cid:28) L (dashed green line). We use typical parameters for the fly
wing: D = 1 µm2/s, DE = 10 µm2/s, α = 0.01 µM/s, αE = 0.001 µM/s, w/L = 0.1,
Cth = 0.1 µM, βE = 10−4/s, β0 = 10−2/s, Eth = 1 µM, L = 25...135 µm (24, 26).
This choice of parameters allows us to illustrate deviations from perfect shape scal-
ing. Other parameter sets might approximate scaling better and might be practically
indistinguishable from the mathematical limit of perfect shape scaling.
The latter can be understood most clearly if we ask what relationship between β and
E would be needed for the steady-state morphogen range λ∗ = χλL to scale with a
constant factor χλ. In Appendix A.2.2, we provide a detailed derivation for different
scenarios. Here, we only discuss the case of a scaling morphogen source (w ∝ L) as a
representative example. If we combine Eq. 2.3, 2.14 and 2.15, we obtain
(cid:18) βE
αEχλ
E∗(cid:19)
λ∗ = χλL ⇒ β∗ ∝ cosh
(2.16)
Thus, this functional relationship between β∗ and E, which would be required to achieve
perfect shape scaling, is different from the relationship, we had originally assumed
in Eq. 2.13. Note the qualitative difference between the two relationships: whereas
Eq. 2.16 corresponds to a positive feedback loop, Eq. 2.13 describes a negative feed-
back loop. We conclude that the considered case of Eq. 2.12 and Eq. 2.13 does not
yield perfect shape scaling, though approximate scaling is nonetheless possible. Such
approximate scaling can be practically indistinguishable from perfect shape scaling as
has been demonstrated by numerical simulations of the discussed feedback and related
versions (24, 26, 135).
31
(a)(b)PositionMorphogen concentrationmorphogen sourceexpandersourcethreshold0Relative position0 0.250.5 0.751 Relative concentration00.20.40.60.81(c)System size050100150Gradient range 101520252. SCALING IN MORPHOGEN SYSTEMS
2.4.4. Other feedback schemes for self-organized gradient scaling
The classical expansion-repression mechanism relies on a feedback in which the mor-
phogen suppresses the expander while the expander increases the range of the mor-
phogen. In consequence, this forms an overall negative feedback loop. In Sec. 2.4.3
we discussed an example which does robustly lead to approximate scaling, though not
perfect shape scaling. Would perfect shape scaling be possible if we deviate from the
scheme of a negative feedback loop?
Eq. 2.16 provides the functional relationship between β and E that would be required
for perfect shape scaling in the steady state. Thus, we could replace Eq. 2.13 by
β ∝ cosh (E/Eth) .
(2.17)
Then, the system possesses a steady state pattern that scales perfectly by construction.
However, this steady state is unstable in numerical simulations as it forms an overall
positive feedback loop. An excess in expander would reduce the level of morphogen by
enhanced degradation and thus the expander production and consequently the expander
level would increase even further. Depending on the initial conditions, the system
shows two distinct dynamics. For small expander levels below the unstable steady
state, the expander source will eventually be suppressed completely and the morphogen
concentration will diverge. For large expander levels, the system will converge to a
second, non-scaling steady state.
Variations of the expander feedback show qualitatively similar results. -- In Ap-
pendix A.2.2-A.2.3, we discuss further variations of the feedback scheme of Eqs. 2.12-
2.13. For these variations, we likewise find approximate scaling, though no perfect shape
scaling in a strict mathematical sense over a large size range. As a general rule, the
ideal case of perfect shape scaling and stability are not easily teamed up. We strongly
emphasize that this does not imply that the mechanisms for approximate scaling can-
not provide suitable theoretical descriptions for gradient scaling in biological systems.
Furthermore, more complex feedback schemes might be able to achieve perfect shape
scaling or at least closely mimic perfect scaling over a considerable size range (24).
2.4.5. Conclusions from our analysis of expander models
Above, we discussed simple realizations of scaling mechanisms for morphogen profiles,
representing two main classes of such mechanisms, see Fig. 2.5. If the expander mea-
sures system size independently of the morphogen, this can result in a robust size
32
2.4 Scaling of morphogen profiles in pre-patterned systems
Figure 2.5.: Various mechanisms have been proposed for the scaling of morphogen
profiles using an expander as a chemical size-reporter: (a) the first class of models
considers an expander which reads out system size independently of the morphogen,
see (78, 90, 96, 135, 148, 154, 233, 234) for examples, (b) the second class of models
assumes a feedback loop between the morphogen and the expander such that the mor-
phogen profile scales itself (24, 26). We demonstrated that in this case it is challenging
to obtain a perfectly scaling steady state. The table summarizes the results for a mor-
phogen that acts on the expander production like in Eq. 2.12, see Appendix A.2.2-A.2.3
for details. For all cases, it is possible to construct a robust feedback yielding good
approximations to the idealized case perfect shape scaling.
read-out and potentially lead to scaling (78, 90, 96, 135, 148, 154, 233, 234). Yet, it re-
quires a tightly controlled layout of the system, e.g. pre-patterned sources and sinks for
the expander. Thus, it does not easily comply with the ability to regenerate and might
just shift the problem from the scaling of the morphogen gradient to the establishment
of a source and sink pattern for the expander.
The second class of models aims to establish scaling of the morphogen profile in a self-
organized manner by a feedback loop between the morphogen and the expander (24, 26).
It turns out that it is challenging to find mechanisms with steady state patterns that
are both: perfectly scaling in the sense of Eq. 2.6 and stable. If a feedback is chosen,
such that a perfectly scaling steady state exists like in Eq. 2.17, this fixed point might
not to be stable. Still, many models like the one defined by Eqs. 2.12-2.13 yield stable
steady states, which can well approximate perfect scaling, depending on the choice of
parameters (24, 26).
33
perfectly scaling steady state is stable in a finite size rangeIndependently controlled expander (e.g. expander-dilution model)Expander feedback loop(e.g. expansion-repression model)- stable scaling(if sources and sinks of the expander are well controlled)degradation ratediffusion coefficientadjustment ofperfectly scaling steady state is unstableno perfectly scaling steady statescaling morphogen sourceconstant morphogen source(a)(b)- approximate scaling is well achievable, perfect shape scaling is challenging- for the simple model variants discussed here, we obtain:perfectly scaling steady state is stable in a finite size range2. SCALING IN MORPHOGEN SYSTEMS
The intuitive picture is that an overall negative feedback loop between morphogen and
expander ensures a stable steady state. In contrast, two mutually suppressing molecular
species, which might yield perfect shape scaling, form an overall positive feedback loop.
Thus, they will not result in a stable steady state unless there are additional feedback
loops. In Chapter 3, we discuss a novel self-organizing scaling mechanism that builds
on such feedback regulations.
An important signature of the second class of models is that the morphogen amplitude
varies with size L. This is a direct consequence of the feedback with a quickly spreading
expander, which ensures a uniform expansion of the gradient range across the system. It
can be understood intuitively because the morphogen not only couples to the expander
but also the expander couples to the morphogen, hence, both have to encode the system
size. This can result in constraints on the system design. For example, it led to the
observation that the combination of a scaling morphogen source and an adjustment of
morphogen diffusion by the expander could not yield scaling of the morphogen profile.
We will have to take this into account when discussing scaling in self-organized systems
in Chapter 3.
2.5. Revisiting the absence of scaling in Turing patterns
In this thesis, we consider one specific choice of a Turing system, which is particularly
suitable for analytical treatment. We restrict our analysis to a one-dimensional system
with reflecting boundary conditions. The dynamics of the two molecules are described
by
∂tA = αA P (A, B) − βA A + DA ∂2
x A
∂tB = αB P (A, B) − βB B + DB ∂2
x B .
(2.18)
Thus, we specifically consider linear degradation with rates βA, βB and production
with rates αA, αB. The function P (A, B) implements a switch-like response to the
concentrations A and B. Production is switched on if the activator concentration
exceeds the inhibitor concentration, A (cid:29) B. We choose a Hill function, which arises
naturally from cooperative and competitive chemical reactions in biological systems,
see Appendix A.4:
(2.19)
In the limit h → ∞, the Hill function can be replaced by the Heaviside theta function
Ah + Bh .
P (A, B) =
Ah
P (A, B) = Θ(A − B) .
(2.20)
34
2.5 Revisiting the absence of scaling in Turing patterns
As a technical point, we have to define Θ(0) != 0, in order for the homogeneous steady
state to always exist, see Appendix A.5. A Fermi function could be another choice to
account for switch-like production, yielding similar results.
In the case of the production switch in Eq. 2.20, the relative source size is defined by
(cid:90) L
0
(cid:96)/L = (cid:104)P(cid:105) =
1
L
P dx ,
(2.21)
where brackets denote a spatial average over the system. We use the same quantity
also to define the source size for the Hill-type production function. This is especially
well justified if the Hill exponent h in Eq. 2.19 is large.
Next, we analyze our example of a Turing system given by Eq. 2.18 and demonstrate
the absence of scaling. We derive the full hierarchy of steady state patterns and discuss
their existence and stability, thereby going significantly beyond linear stability analysis
of the homogeneous steady state. This will put us in the position to understand the
mechanisms for scaling of a Turing system.
2.5.1. A hierarchy of steady state solutions
Eq. 2.18 together with the production function in Eq. 2.19 possesses a unique homoge-
neous steady state:
A∗
h =
αA/βA
1 + (βAαB/(αAβB))h
, B∗
h =
For h → ∞, the steady state concentrations approach zero, A∗
pendix A.5.
The inhomogeneous steady states can be computed analytically in the limit of h → ∞,
corresponding to the source switch of Eq. 2.20, see Appendix A.6. We find a hierarchy
h = 0, see Ap-
αB/βB
1 + (βAαB/(αAβB))h .
h = B∗
(2.22)
of steady state patterns, which can be characterized by the number m of contiguous
source regions with A > B and the number n of source regions touching the system
boundaries, see Fig. 2.6. Any (m, n)-pattern can be constructed as a concatenation of
σ = 2m−n copies of the (1, 1)-pattern as a basic building block, see Appendix A.6. Note
that s = σ/2 is the corresponding leading order Fourier mode of the spatial patterns.
The (1, 1)-pattern with a single source, located in the interval 0 ≤ x < (cid:96), is given by
sinh(L/λA)
1 − sinh(L/λA−(cid:96)/λA)
(cid:16) x−L
1 − sinh(L/λB−(cid:96)/λB)
(cid:16) x−L
sinh((cid:96)/λA)
sinh(L/λA) cosh
sinh((cid:96)/λB)
sinh(L/λB) cosh
sinh(L/λB)
λB
λA
(cid:17)
(cid:17)
cosh
cosh
(cid:17)
(cid:17)
(cid:16) x
(cid:16) x
λA
λB
A∗
(1,1) =
B∗
(1,1) =
αA
βA
αB
βB
, 0 ≤ x ≤ (cid:96)
, (cid:96) < x ≤ L
, 0 ≤ x ≤ (cid:96)
, (cid:96) < x ≤ L .
(2.23)
(2.24)
35
2. SCALING IN MORPHOGEN SYSTEMS
Figure 2.6.: (a) The steady state solutions of Eq. 2.18 with Eq. 2.20 can be char-
acterized by a pair of pattern numbers (m, n), where m is the number of contiguous
source regions (gray) with A > B and n is the number of source regions touching
the system boundaries. (b) Any (m, n)-pattern can be constructed as a concatenation
of σ = 2m − n copies of the (1, 1)-pattern as a basic building block, here shown for
(m, n) = (2, 2). (c) Thus, we obtain a complete hierarchy of steady state solutions,
exemplified for the concentration A up to m = 3 sources. Parameters: DB/DA = 30,
αB/αA = 4, βB/βA = 2, λA/L =
0.1 ≈ 0.3, λB/L =
1.5 ≈ 1.2.
√
√
Patterns are characterized by two pattern length scales for A and B, respectively,
analogous to Eq. 2.5.:
λA =(cid:112)DA/βA and λB =(cid:112)DB/βB .
The source size (cid:96) has to satisfy the implicit equation
αA βB
βA αB
sinh((cid:96)/λA)
sinh(L/λA)
cosh
=
sinh((cid:96)/λB)
sinh(L/λB)
cosh
(cid:18) (cid:96) − L
(cid:19)
λA
(cid:18) (cid:96) − L
(cid:19)
λB
(2.25)
(2.26)
[0.1cm] Fig. 2.7 depicts the relative source size (cid:96)/L, which is a function of only three
dimensionless ratios: L/λA, L/λB and αAβB/(βAαB), see also Appendix A.6. A dimen-
sional analysis of Eq. 2.18 shows that this is true also for finite h and other production
36
(a)(b)=+(c)source2.5 Revisiting the absence of scaling in Turing patterns
Figure 2.7.: Pattern length scales set size of source: The relative source size (cid:96)/L of
the (1, 1)-pattern at steady state is uniquely determined by the pattern length scales
L/λA and L/λB.
If only the system size is varied, while the other parameters are
kept constant, the relative source size changes along the red curve (here: λB/λA = 5).
Following the red curve, the source size approaches (cid:96)/L ∝ λB/L for λB (cid:28) L. Fur-
thermore, we obtain a threshold for the existence of the (1, 1)-pattern (green) and a
scaling regime in the limit of λB (cid:29) L (cid:29) λA (saturation of blue curve). Parameter:
αAβB/(βAαB) = 0.5.
functions. The red line illustrates the variation of the relative source size (cid:96)/L as a
function of L, when all other parameters are kept constant. Note that an identical
curve could be obtained if instead of L, the length scales λA and λB would be varied
via a common control parameter. This will become important in Chapter 3 when we
combine the Turing system with an expander feedback.
The source size is zero in systems that are very small relative to the characteristic
length scales, see Fig. 2.7 (green curve). For these small systems, the homogeneous
steady state is the only existing solution. The (1, 1)-pattern begins to appear only
above a critical size L1, where L1 obeys
βA αB
αA βB
tanh(L1/λA)
L1/λA
=
tanh(L1/λB)
L1/λB
.
(2.27)
This equation implies in particular that λA has to be sufficiently small in comparison
to L and λB for the existence of inhomogeneous patterns, see Appendix A.6.
37
00.12551015210300.25changing, whileconstantthreshold for:limit of:2. SCALING IN MORPHOGEN SYSTEMS
The relative source size (cid:96)/L strongly increases with L beyond L1.
In the limit of
λB (cid:29) L, it eventually saturates, see Fig. 2.7 (blue curve). This observation of source
scaling for an existing (1, 1)-pattern has been reported before (66, 67, 120). In contrast,
following the red curve for a finite inhibitor range λB, the source size (cid:96) approaches a
constant value independent of L. In the limit λB (cid:28) L, we obtain (cid:96)/L ∝ λB/L. The
source does not sense the system boundaries but instead is restricted by λB.
Analogous to the (1, 1)-pattern, any higher order (m, n)-pattern only exists if the system
size L exceeds a critical size Lσ = σ L1. Fig. 2.8(a) shows the region of existence (gray)
for patterns of different pattern numbers.
2.5.2. Higher order patterns form in larger systems
We have shown that the number of coexisting patterns increases with increasing size.
Next, we investigate which of these patterns are stable and can in fact be observed in
systems of different sizes.
Higher order modes become linearly unstable in larger systems. -- The homogeneous
steady state is linearly unstable with respect to inhomogeneous perturbations for pa-
rameter values that fulfill the Turing conditions discussed in Section 1.4. A linear
stability analysis as detailed in Appendix A.5 provides the corresponding parameter
range for our specific choice of Eq. 2.18 and Eq. 2.19. For a certain choice of parame-
ters and a specific system size, there is a range of perturbation modes s ∈ [smin, smax],
which grow exponentially and might result in inhomogeneous patterns. Importantly,
the mode number s scales with system size L as can already be seen from the struc-
ture of the general linear stability matrix in Eq. 1.4. Thus, both interval bounds smin
and smax linearly depend on L as does the mode that becomes unstable first, when
changing one control parameter. Therefore, in a larger system, more and higher order
perturbation modes become unstable (64, 96, 148).
Nevertheless, the linear instability of the homogeneous steady state is only a necessary
condition for the formation of inhomogeneous patterns. The mode number s of the per-
turbation can provide a first clue about the type of pattern that might form, yet there
is no general relationship between the growing perturbation modes and final patterns.
In fact, it is not even clear that an inhomogeneous steady state exists in a system of
finite size, even if the homogeneous steady state is linearly unstable. We encounter
such a situation in our system.
38
2.5 Revisiting the absence of scaling in Turing patterns
Figure 2.8.: Classical Turing patterning implies that in larger systems higher-order
patterns form. (a) Typical profiles of the activator concentration A(m,n)(x) for the
(m, n)-pattern are shown in red as steady state solutions to Eq. 2.18. Size ranges are
shown, where the (m, n)-pattern is linearly stable (black), or exists, but is not stable
(gray). In the blue region, the (1, 1)-pattern is the only stable pattern. (b) Basins
of attraction: final pattern type at steady state as a function of system size on the
horizontal axis and initial conditions on the vertical axis. Initial conditions linearly
interpolate between the (1,1)- and (1,0)-pattern, i.e., A(x, t=0) = (1 − ξ) A(1,1)(x) +
ξ A(1,0)(x), and analogously for B(x, t=0). Parameters: DB/DA = 30, αB/αA = 4,
βB/βA = 2, h→∞ (a), h = 5 (b).
39
0102030(a)pattern numbers(0, 0)(1, 1)(1, 0)(2, 2)(2, 1)100.50102030System sizeInitial conditions(b)2. SCALING IN MORPHOGEN SYSTEMS
The Turing system possesses a rich dynamics including oscillations due to a finite
system size and the multistability of several steady states. -- Fig. 2.8(a) summarizes
size ranges of existence (gray) and stability (black) of steady state patterns of our model.
For very small systems with L < L1, the homogeneous steady state is the only steady
state pattern that exists in agreement with reflecting (and also periodic) boundary
conditions. Still, the homogeneous steady state can be linearly unstable. As there are
no diverging terms in Eq. 2.18, we observe homoclinic orbits and oscillations. These
oscillations induced by boundary effects take place far away from the homogeneous
steady state. Thus, they appear even if all eigenvalues of the linear perturbation matrix
Ms are real.
For larger systems with L > L2, several inhomogeneous steady states coexist and
are linearly stable. We numerically determined the region of linear stability (black)
for several inhomogeneous steady states, see Appendix A.6.4. We did not find any
upper size limit for the linear stability for a particular pattern. However, we observed
increasingly smaller basins-of-attraction in systems of increasing size. Thus, lower order
patterns become unstable with respect to finite-amplitude perturbations in favor of
higher wavenumber patterns, as exemplified in Fig. 2.8(b).
In conclusion, steady state patterns of increasing wavenumber emerge with increasing
system size. -- In particular, the interplay between the characteristic length scales λA
and λB and the system size L determines whether a certain pattern fits into the system
and whether it will preferentially form. For example, the (1, 1)-pattern is globally
stable only in a limited size range, see Fig. 2.8(a) (blue shading). Therefore, only if the
system is pinned to this blue region by changing the characteristic length scales with
system size, the (1, 1)-pattern emerges reliably from arbitrary initial conditions. In a
hypothetical case, in which λA and λB scale with L, while αAβB/(βAαB) stays constant,
we can hope to find a range of parameters, for which a particular (1, 1)-pattern with
a scaling source is the only stable steady state, see Eq. 2.26 and Fig. 2.7. In the next
Chapter, we introduce a corresponding feedback that couples the Turing system to the
dynamics of an expander molecule.
40
2.6 Summary
2.6. Summary
In this chapter we analyzed scaling of morphogen profiles in pre-patterned and self-
organized systems. For this, we especially paid attention to the difference between
(perfect) scaling of the profiles and approximate scaling.
For the case of pre-patterned systems, several scaling mechanisms have been proposed,
drawing on the idea of additional expander molecules, which encode system size (24, 26,
135, 233, 234). We distinguish two classes of scaling mechanisms. In the first class, the
expander reads out system size independently of the morphogen like in the expander-
dilution model. The second class comprises mechanisms of self-scaling by a feedback
loop between expander and morphogen like in the expansion-repression model. These
latter models typically yield robust approximate scaling but often do not perfectly
scale. We highlight several requirements of such expander feedbacks, in particular the
fact that the amplitude of the morphogen profile has to vary with system size.
As a prototype example for self-organized patterning, we discussed a classical Turing
system with two players. Our specific choice of equations is especially suitable for
analytical treatment. We illustrated the absence of scaling in a comprehensive way:
Besides the linear stability analysis of the homogeneous steady state, we also assessed
existence and stability of the inhomogeneous patterns. We found a hierarchy of in-
homogeneous steady state patterns, for which higher order patterns are favored with
increasing system size. Thereby, the behavior is governed by the ratio of system size
L to the characteristic length scales λA and λB of the two considered molecules, indi-
cating that scaling might be achieved by adjusting these length scales with L. In fact
similar ideas have already been explored in earlier works (90, 96, 148, 154). It has been
shown that by coupling λA and λB to an autonomous chemical size reporter like in the
expander-dilution model, the Turing patterns scale. In Chapter 3, we demonstrate that
we can also devise an expander feedback loop in the spirit of the expansion-repression
model to generate self-organized scaling of a Turing system.
41
3. Scaling and regeneration
of self-organized patterns
3.1. Self-scaling and self-organization
Flatworms challenge previous theories of pattern scaling and self-organization. Regene-
ration from minute amputation fragments of strongly varying size prompts for patter-
ning mechanisms that are both: scalable and self-organizing. Turing systems, as promi-
nent frameworks for self-organized patterning, do not naturally scale (74, 148, 223).
They typically generate the same patterns repeated multiple times in a growing system
as depicted in Fig. 3.1.
In Chapter 2, we have illustrated that this is due to fixed
characteristic length scales, which define the wave length of the patterns.
However, it has been demonstrated that Turing patterns can in fact scale if the charac-
teristic length scales are coupled to a chemical size-reporter, which we called expander
"Everywhere nature works true to scale, and
everything has its proper size accordingly." --
D'Arcy W. Thompson, On Growth and Form,
1945 (217)
(90, 96, 148, 154). These previous works assumed
an expander that is independently controlled by a
pre-patterned system like in Section 2.4.2. Thus,
the proposed mechanisms account for scaling, but
not in a fully self-organized way. Conversely, self-
scaling mechanisms have been discussed previously (e.g. for fly development), yet for
systems that are not fully self-organized (11, 24, 27, 223, 234), see Section 2.4.3.
Now, in this chapter, we combine the two features of scaling and self-organization and
present a generic mechanism that yields fully self-organized and self-scaling patterns.
Importantly, the expander feedback in a Turing system with a self-organized source
follows a very different control logic than in systems with a pre-existing source. In con-
sequence, perfectly scaling steady states are stable across a wide range of system sizes.
Our novel class of self-scaling Turing patterning provides a conceptual framework to
address scaling and regeneration in flatworms and guides the design and interpretation
of experiments. This result has been published in Werner et al. (239).
First, we present a specific example which combines ideas from Turing patterning, dis-
cussed in Section 2.5, with theoretical concepts for the scaling of morphogen profiles,
42
3.2 A minimal model for self-organized pattern scaling
Figure 3.1.: Classical Turing patterns show more periodic repeats in larger systems
as a result of fixed intrinsic length scales (a), instead of being a scaled-up version of
the patterns in small systems (b).
analyzed in Section 2.4. We demonstrate that this system is capable of self-organized
scaling and explain the underlying mechanism. In a second step, we discuss generaliza-
tions of the mechanism and predictions for experiments in flatworms.
3.2. A minimal model for self-organized pattern scaling
Analogous to Section 2.4, we add another molecular species, termed expander, whose
concentration E will provide a read-out of system size L. As a specific case, we assume
the expander is produced homogeneously, spreads by diffusion and is subject to degra-
dation
∂tE = αE − βE E + DE ∂2
x E .
(3.1)
Again, αE denotes the respective production rate, βE the degradation rate and DE an
effective diffusion coefficient.
In order to achieve self-organized scaling, the dynamics of the three molecular species
have to be mutually coupled in a similar manner as for the expander feedback in
Section 2.4.3. As an illustrative example, we choose a feedback loop, in which the
inhibitor of the Turing system controls the degradation rate of the expander via
βE = κE B
(3.2)
with a positive constant κE. This choice ensures that the expander concentration is
approximately homogeneous even for small DE because the inhibitor concentration
itself is typically rather homogeneous in a Turing system.
In turn, the expander also feeds back on the Turing system and changes the length
scales of the morphogen profiles, see Fig. 3.2(a). Analogously to Eq. 3.2, we choose a
linear regulation of the degradation rates by the expander (with κA, κB > 0)
βA = κA E ,
βB = κB E .
(3.3)
Note that Eqs. 2.18 and 3.1 can be considered as a classical three component Turing
system, yet with specifically chosen coupling terms given by Eqs. 2.19, 3.2 and 3.3.
43
(a) Classical Turing(b) Scaled Patterns3. SCALING & REGENERATION OF SELF-ORGANIZED PATTERNS
3.3. Numeric solution shows scaling and pattern regeneration
Before analyzing this system of equations analytically, we numerically demonstrate that
the steady state pattern scales with system size over several orders of magnitude, see
Fig. 3.2(b)-(c). Small deviations from the perfect collapse of the concentration profiles
arise due to a finite diffusion coefficient of the expander and a finite h of the production
function. Still, the source size (cid:96)∗ at steady state and the characteristic length scales λ∗
A
and λ∗
B show almost perfect scaling with system size. In the limit of an homogeneous
expander and a switch-like source with h → ∞, we analytically show in Section 3.4
that the scaling becomes exact.
Pattern scaling does not come at the cost of robustness for the Turing system. We
can challenge patterning by perturbations that mimic amputation experiments, see
Fig. 3.2(d). Here, the activator concentration A reliably adjusts to fit the smaller
system size. Note that the system part with the initially oversized source and thus
higher morphogen concentrations takes longer to re-adjust.
3.4. Dynamical systems analysis of scaling and regeneration
Next, we provide insight into how and why scaling works. First, we identify steady
states, each of which scales with system size. For the simple case of adiabatically slow
expander dynamics, we then show that the (1,1)-pattern is a stable steady state. A
phase space description allows us to characterize fixed points and understand the non-
linear dynamics. We illustrate the behavior by considering a two-dimensional projection
of the phase space of our dynamical system, spanned by the relative source size given
by Eq. 2.21 and the mean expander level, see Fig. 3.3(a).
An implicit scaling relation holds at steady state. -- We first show that all steady
states of the extended Turing system are characterized by the same relative source
size (cid:96)∗/L = (cid:104)P ∗(cid:105) independent of system size L. Thus, in Fig. 3.3(a) all steady states
(circles) are found on a horizontal line. By spatial averaging of Eq. 2.18 and 3.1, we
obtain for steady state concentrations B∗ and E∗
0 = αB(cid:104)P ∗(cid:105) − κB(cid:104)B∗E∗(cid:105)
0 = αE − κE(cid:104)B∗E∗(cid:105) .
Hence, the relative source size in the steady state is constant
(cid:96) *
L
=
αE κB
αB κE
.
44
(3.4)
(3.5)
(3.6)
3.4 Dynamical systems analysis of scaling and regeneration
Figure 3.2.: Scalable pattern formation in a Turing system with expander feedback.
(a) The Turing system and the expander mutually control their degradation rates, re-
sulting in a stable feedback loop. (b) Scaling corresponds to morphogen profiles that
collapse as a function of relative position x/L (normalized by respective concentrations
A0, B0, E0 at x = 0). (c) The feedback self-consistently adjusts the length scales λA
and λB of the morphogen profiles and thus the source size (cid:96) with system size (sym-
bols: numerical results; lines: analytical solution of Eqs. 2.18 and 3.1 at steady state
√
for homogeneous expander concentration and h→∞). Here, λ0 = [DA/
αAκA]1/2
denotes the characteristic length scale of the system. (d) Regeneration of concentra-
tion profiles: after cutting the system in two (initial size L/λ0 = 20), the expander
feedback adjusts the activator concentration to fit the smaller system size both in a
tail fragment (upper row) and head fragment (lower row). Repatterning of the tail
fragment takes significantly longer due to the initially higher morphogen concentration
√
αAκA). Parameters: DB/DA = 30, DE/DA = 10,
(time given in units of τ0 = 1/
αB/αA = 4, αE/αA = 0.4, κB/κA = 2, κE/κA = 2, h = 5.
45
(a)10−110010110210010110200.511.500.51(c)Pattern length scalesRescaled concentration profilesRelative positionSystem size(b)(d)Activator profile001224366060120180240Activator profile3. SCALING & REGENERATION OF SELF-ORGANIZED PATTERNS
In contrast, the length scales defined in Eq. 2.25 are not constant anymore but change
(cid:114) DA
κAE
A and λ∗
(cid:114) DB
κBE
with expander level:
λA =
and λB =
.
(3.7)
We have seen in Fig. 3.2(c) that λ∗
with system size.
B at steady state scale with high precision
In order to explain this scaling behavior, we consider the limit
of a spatially homogeneous expander concentration E, for which we can show that
the scaling becomes exact. This limit corresponds to either a large expander range
λE =(cid:112)DE/(κEB) (cid:29) L or a large inhibitor range λB =(cid:112)DB/(κBE) (cid:29) L. The latter
condition is in compliance with the requirements for a typical Turing system and is
promoted by the scaling feedback.
If the expander level was imposed as constant E = E0, we would obtain a Turing system
without expander feedback, as discussed in Section 2.5. Thus, the concentrations would
converge towards one of the (m, n)-patterns discussed in Section 2.5.1 with pattern
length scales λA(E0) and λB(E0). The relative source size f(m,n) = l/L of such a
pattern depends on E0 only via the dimensionless ratios λA(E0)/L and λB(E0)/L, see
Eq. 2.26. This fact is illustrated by Fig. 2.7, where the red curve displays f(1,1) in the
limit of h → ∞. Importantly, f(m,n) = f(m,n)(L2E0) is only a function of L2E0 and
changing E0 has analogous effects on the relative source size as changing L2.
The same argument also implies that a (m, n)-pattern can only exist above a critical
value of E0, corresponding to the minimum system size for the existence of patterns in
Fig. 2.8(a). Below this critical value, f(m,n) is zero. Above this value, f(m,n) displays
a nonmonotonic dependence on E0, which results from opposing effects of the pattern
length scales of the activator and the inhibitor on the source size (cid:96), see Fig. 3.3(a).
The intersections of the curves f(m,n) (obtained from the steady states of the Turing sys-
tem without expander) with the constant value (cid:96)∗/L given by Eq. 3.6 define the steady
states of the full system with expander feedback. For each pattern type (m, n), we
find two steady-state patterns, denoted (m, n)+ and (m, n)−, with respective expander
levels E+
The fact that f(m,n)(L2 E∗) = (cid:96)∗/L is independent of system size L by Eq. 3.6, implies
that also L2E∗ is independent of L for each steady state. We conclude E∗ ∝ L−2 and
B ∝ L at each fixed point, consistent with our numerical results in
thus λ∗
Fig. 3.2(c).
(m,n), see the black and white circles in Fig. 3.3(a).
(m,n) < E−
A ∝ L, λ∗
The stability in the vicinity of the scaling fixed points can be assessed in the simple limit
of adiabatically slow expander feedback. -- In this limit, the source size first relaxes to
46
3.4 Dynamical systems analysis of scaling and regeneration
Figure 3.3.: Fixed points and dynamics in a Turing system with expander feedback.
(a)-(b) Two-dimensional projection of the phase space: all fixed points (circles) are
characterized by the same relative source size. For adiabatically slow expander dy-
namics, the system relaxes along the nullclines of the Turing system f(m,n) (shown for
h → ∞, λE (cid:29) L, corresponding to Eq. 2.26). As each nullcline intersects the steady-
state condition of Eq. 3.6 twice, the system possesses two fixed points (n, m)+ and
(n, m)− for each pair (n, m). In the blue region, the (1,1)-pattern is the only stable
steady state of the Turing system, compare to Fig. 2.8, implying that all trajectories
starting there must converge to this fixed point. Here, E0 = (αA/κA)1/2 denotes a
characteristic concentration of the system. (c) Example trajectories, mimicking am-
putation experiments (labeled i,ii ), and uniform, one-time injection of the expander
(labeled iii,iv ); all converge to the same stable fixed point, an appropriately scaled
(1,1)-pattern. (d) Parameter regions for stable, self-scaling pattern formation (green),
and regions of expander divergence (orange, purple). Parameters of panel (c) and
Fig. 3.2 indicated by cross. Parameters: DB/DA = 30, DE/DA = 10, αB/αA =
4, αE/αA = 0.4,
h = 5, L/λ0 = 10, unless indicated
otherwise.
κB/κA = 2,
κE/κA = 2,
47
10−110000.10.2(c)10−110000.10.2(a)(b)025500123Relative source sizeRelative source sizeMean expander concentrationMean expander concentrationstable scaling(1,1)+(d)(1,1)−(1,0)+(1,0)−(2,2)−(2,2)+(1,1)(2,2)(1,1)+(1,1)−(1,0)+(2,2)+(1,0)−(2,2)−3. SCALING & REGENERATION OF SELF-ORGANIZED PATTERNS
(cid:96)/L = f(m,n)(L2 E) for some (m, n), corresponding to the fast time scale of the Turing
system. Then, by Eq. 3.1, the system moves slowly along this nullcline according to
∂tE = αE − κE αB
κB
f(m,n)(L2 E) .
(3.8)
Stability of steady-state patterns requires ∂Ef(m,n) > 0, which is observed to hold only
for E+
(m,n), see Fig. 3.3(a).
Which branch f(m,n) is selected for arbitrary initial conditions by the fast Turing dy-
namics? As we consider the limit of slow expander feedback, the expander concentra-
tion is approximately constant within the time scale of the Turing relaxation. Thus,
the question of which pattern (m, n) is selected is formally equivalent to the stability
of (m, n)-patterns in the Turing system without expander feedback as a function of
system size L. The blue region in Fig. 2.8(a), in which the (1, 1)-pattern is the only
stable pattern, can be associated with the blue range of expander concentration in
Fig. 3.3(a). Therefore, we can deduce that this represents a basin-of-attraction for the
(1, 1)+-pattern.
A numerical analysis reveals a large basin-of-attraction for the first order pattern. --
Fig. 3.3(c) shows that the (1, 1)+-pattern is an attractive fixed point also for trajectories
starting outside this blue region and for nonadiabatic expander dynamics. Two example
trajectories, labelled (i ) and (ii ), corresponding to head and tail fragments, respectively,
converge to the (1, 1)-pattern, after a transient under- and over-shoot of the source size.
Two additional trajectories, labeled (iii) and (iv ), simulating uniform injection of the
expander, likewise converge to this fixed point. We can understand these observations
by drawing on the results for the Turing system without expander in Section 2.5.1 and
2.5.2.
The dynamics for large expander concentrations are characterized by a successive move-
ment through higher order patterns. -- The (1, 1)+-state is also attractive for trajec-
tories like (iv) with initially high values of E. Here, a (2, 2)-pattern with an additional
source is transiently emerging. Yet, according to the expander dynamics given by
Eq. 3.8, the system moves towards the fixed point (2, 2)+ as long as the (2, 2)-pattern
comprises a stable fixed point of the Turing system without expander. However, the
fixed point (2, 2)+ of the Turing system with expander lies in a region of the phase
space where the (2, 2)-pattern of the Turing system alone is unstable for our choice
of parameters. Thus, the fixed point (2, 2)+ is in fact a saddle point, attractive with
respect to the expander dynamics along the nullcline f(2,2) but unstable with respect
to the Turing system. Again, this can be understood from Fig. 2.8(a) by noting the
48
3.4 Dynamical systems analysis of scaling and regeneration
formal correspondence between changes in E for a fixed system size L and changes in
L2 for the classical Turing system without expander. Fig. 2.8(a) illustrates that the
(2, 2)-pattern is unstable outside the blue region. The result would be different for the
(1, 0)-pattern. Note that the nullclines of the (1, 0)-pattern and the (2, 2)-pattern are
congruent in our projection of the phase space. In consequence, there exists a trajectory
starting with a (1, 0)-pattern that looks identical to (iv), besides the fact that it ends
in the fixed point (1, 0)+, which is stable for our choice of parameters, as illustrated in
Fig. 2.8(a).
Dynamics of homogeneous states allows to understand regeneration of patterns. --
For small expander values, there are no inhomogeneous steady state patterns of the
Turing system, yet homogeneous patterns exist. To understand the relaxation of the
full system, we will discuss two limiting cases, assuming again the separation of time
scales between Turing dynamics and expander system:
a) In the limit of fast Turing dynamics, the Turing system will first approach the
homogeneous steady state, given by Eq. 2.22 for a certain value of E. On the larger time
scale of the expander relaxation, the system moves along the corresponding nullcline
(cid:96)/L = f(0,0) according to
∂tE = αE − αBκE
κB
f(0,0) with f(0,0) = P (A∗
h, B∗
h) .
(3.9)
If f(0,0) < (cid:96)∗/L, the expander level increases and consequently the Turing length scale
decreases until the homogeneous pattern becomes unstable and inhomogeneous patterns
can form.
b) As a second case, we consider the opposite limit, for which the Turing dynamics are
adiabatically slow and the inhibitor concentration B does hardly change within the time
scale of relaxation of the expander towards αE/(κE B). We find a qualitatively similar
behavior. It exists another homogeneous, yet dynamic solution of the Turing system,
in which the ratio χ = B/A and thus P (A, B) = g(χ) is approximately constant. The
ratio χ is determined by the following implicit relation, see Appendix A.7.1:
1
1 + χh =
αE
κE
κA − κB
αA χ − αB
.
(3.10)
In order for E to increase such that the system enters the regime where inhomogeneous
patterns exist, B has to decrease. According to
∂tB = αB g(χ) − κB αE
κE
,
(3.11)
this requires that g(χ) < (cid:96)∗/L.
49
3. SCALING & REGENERATION OF SELF-ORGANIZED PATTERNS
Depending on the choice of parameters, both transient homogeneous states (corre-
sponding to the two limiting cases of a separation of time scales) can be observed when
mimicking regeneration experiments in the Turing system with expander. In Fig. 3.3(c),
the horizontal stretch of trajectory (i) corresponds to g(χ). Both production terms f(0,0)
and g(χ) of the transient homogeneous states have to be smaller than (cid:96)∗/L, explaining
the transient undershoot during regeneration. Note that from g(χ) = (cid:96)∗/L follows that
also f(0,0) = (cid:96)∗/L. The set of parameters corresponding to this joint condition mark the
breakdown of robust regeneration, see horizontal line in Fig. 3.3(d). Beyond this thres-
hold, the homogeneous morphogen concentrations diverge while the expander vanishes
(purple region). If the parameters obey this constraint, the basin-of-attraction of the
(1, 1)-state extends towards small values of E outside the blue region in Fig. 3.3(a).
3.5. As single source pattern is attractive for a large region of
the phase space
Based on the discussion above, we expect the (1, 1)-pattern to be re-established even
when simulating more drastic amputation experiments, as shown in Fig. 3.4. Again,
we find the source (cid:96)/L of each smaller system first approaching the horizontal line
corresponding to g(χ). Eventually, the expander value is sufficiently large such that
the homogeneous state becomes unstable and inhomogeneous patterns can form. Tran-
siently, the (2, 1)-pattern arises but as the (2, 1)+-solution is unstable with respect to
the Turing dynamics, the additional source vanishes and the system returns to the
(1, 1)-pattern.
Details of the relaxation dynamics and whether (1, 1)+ is the only stable fixed point or
whether a second stable fixed point (1, 0)+ with a smaller basin-of-attraction exists like
in our example, depends on the choice of parameters. In general, we observe robust
pattern scaling for a vast parameter range, provided (i) inhibitor diffusion is sufficiently
fast (a necessary condition for pattern formation in any Turing system) and (ii) the
feedback strength is similar for both Turing molecules, see Fig. 3.3(d). The threshold
above which the expander concentration vanishes, resulting in a divergence of A and B
(purple), can be analytically determined from the condition f(0,0) = g(χ) = (cid:96)∗/L. This
corresponds to the fact that the expander should increase for homogeneous morphogen
concentrations.
50
3.6 Structural robustness for pattern scaling
Figure 3.4.: Numerical solution of the Turing system with expander mimicking an
amputation experiment with 10 cuts. All trajectories converge to the fixed point (1, 1)+
after a transient formation of the (2, 1)-pattern. Parameters: DB/DA = 30, DE/DA =
10, αB/αA = 4, αE/αA = 0.4, κB/κA = 2, κE/κA = 2, h = 5, L/λ0 = 10.
3.6. Structural robustness for pattern scaling
Several generalizations of the proposed minimal mechanism are conceivable. First, the
feedback of the Turing system on the expander level could be likewise implemented via
the production rate, e.g. αE ∝ B, instead via the degradation rate βE = κEB. Then,
scaling would require βA ∝ 1/E , βB ∝ 1/E as well as a fast spreading expander,
yielding analogous steady states.
As a second possibility for pattern scaling, the feedback in Eq. 3.3 could also be medi-
ated by A (instead of B or additional to B), provided the expander diffuses sufficiently
fast. For example, the following system can be interpreted in the sense that pairs of A
51
10−310−210−110000.10.2Relative source sizeMean expander concentration(2,1)+(1,1)+(1,0)+(2,2)+(1,1)−3. SCALING & REGENERATION OF SELF-ORGANIZED PATTERNS
and E as well as B and E are only degraded together when bound to each other
∂tA = αA P (A, B) − κA AE + DA ∂2
x A
∂tB = αB P (A, B) − κB EB + DB ∂2
x B
∂tE = αE − κA AE − κB BE + DE ∂2
x E .
(3.12)
More generally, the degradation term does not even have to be linear in A, B and E.
Two increasing functions of A and E, and B and E, respectively, that fulfill the Turing
conditions together with P (A, B) are sufficient to generate robust scaling by the same
arguments as above, see Appendix A.7.2.
Remarkably, the feedback loop considered here, featuring two mutually suppressing
concentration profiles (B and E), would be unstable for a pre-patterned morphogen
source of fixed size as discussed for the expander feedback in Section 2.4.3. The feedback
mechanism becomes stable, only due to the self-stabilizing effect of the Turing system
itself.
Our mechanism yields pattern scaling by scaling of the morphogen profiles via a feed-
back on the degradation rates of A and B. As discussed in Section 2.3, scaling of expo-
nential profiles can in principle also be achieved by mechanisms affecting the diffusion of
morphogens. However, controlling only diffusion is not compatible with self-organized
pattern scaling as presented here. Our mechanism relies on a size-dependent amplitude
of the morphogen profiles, which is lacking for pure diffusion control, as discussed in
Section 2.4.5.
The morphogen amplitude in our feedback scheme changes quadratically with system
size as can be seen from Eq. 2.23-2.24. It is interesting to note that the flux βAA has
a size-independent amplitude. The spatial profile of this flux could provide a read-out
of the scaling morphogen profiles independent of their amplitudes, see Appendix A.7.3.
For example, if degradation only happens by internalization upon binding to a receptor,
the concentration of down-stream targets of the morphogen A would be proportional
to the flux βAA.
52
3.7 Signatures of self-scaling patterns
3.7. Signatures of self-scaling patterns
Our minimal model for the scaling of self-organized patterns can act as a framework
to understand scaling and regeneration in flatworms. We find a number of signatures
characterizing this class of patterning mechanism, which can be identified in experi-
ments.
(i) First, the over-all morphogen levels depend on system size. As already discussed
in Section 2.4.5 for self-scaling feedback loops, this is necessary for the morphogens
to influence the expander in a size-dependent manner. However, downstream targets
might scale with a constant amplitude.
(ii) Second, for the simple scalable Turing systems discussed here, a size-dependent
morphogen amplitude requires the adjustment of degradation rates. For the experi-
ments, we have to take into account that many processes can effectively lead to a
change in degradation by expander molecules. For example, binding of the expander
could disintegrate or alter the morphogen such that it cannot fulfill its signaling task.
Alternatively, the expander could act as a co-receptor that facilitates internalization
and thus removes the morphogen from the system. Conversely, the expander might
reduce degradation by preventing disintegration, modifications or internalization by
binding to the morphogen or its suppressors.
(iii) Third, the expander by definition shows a size-dependent concentration.
If the
expander level is proportionally to L2, it can couple linearly to the degradation rate
to achieve scaling. Note that the expander is likely not a single molecule but an entire
signaling module, for which some components might already be expressed in a size-
dependent manner. Thus, analyzing changes in expression levels between animals of
different size can help to identify the expander mechanism.
(iv) Fourth, a perturbation of the expander feedback by over-expression or gene knock-
out or also by hindered spreading is expected to lead to a change in the wave number of
the pattern. Depending on the coupling of the expander to the morphogen dynamics,
one might observe the most severe effects in either small or large animals.
(v) Fifth, the source is expected to exhibit a non-monotonic dynamics after ampu-
tation. In particular, an initially oversized source might show an undershoot during
regeneration.
(vi) Sixth, there might be additional fixed points of the system with small basins-of-
attraction. In our example, the (1, 0)+-pattern was also stable and locally attractive.
If the system is initialized close to this pattern, the pattern will be maintained, even in
the presence of small perturbations.
53
3. SCALING & REGENERATION OF SELF-ORGANIZED PATTERNS
3.8. Comparison to experiments in flatworms
The signatures discussed in Section 3.7 can to some extent be found in flatworms, in
particular for the Wnt/β-catenin pathway associated with head-tail polarity (4, 9, 51,
54, 81, 82, 141, 160). The Wnt molecules are expressed in the posterior part of the
animal and shape the graded β-catenin profile with a maximum at the tail, see Fig. 3.5.
The expression regions of components of the Wnt pathway as well as independent
anterior and posterior markers fit to the size of the worm (81). Furthermore, first
preliminary data by our experimental collaborators suggests that also the β-catenin
profile matches worm size with a constant amplitude, see (i).
After amputation, this graded cue has to be quickly scaled to fit the remaining body
fragment. Indeed, it has been observed that the position-dependent expression pattern
along the AP axis changes within the existing cells to adjust to the new fragment size
in the course of a few days (81, 82, 245). Thus, the robust re-establishment of a scaled
chemical pattern seems to predate the remodelling of the body plan. Thereby, the
expression region of one of the Wnt molecules is reported to show similar dynamics as
the source size in our model, see (v). After cutting the worm, the expression of Wnt11.5
is strongly reduced and undershoots its target range before it eventually scales up again
to fit the fragment size (81).
Knockdown of β-catenin leads to the formation of multiple heads (3, 51, 82, 82, 160).
Interestingly, after a transient drug treatment, the double-headed phenotype persists
even after consecutive amputations (117). This agrees with our discussed example for
which one second order pattern (and even more specifically the (1, 0)-pattern) seems to
be a locally attractive fixed point, see (vi).
Additionally, very narrow slices of the worm can also sometimes result in a double-
headed phenotype (4, 51, 117, 127). Again, this compares to our model, for which
the morphogen profile transiently flattens after amputation. In the model, the profile
becomes the more homogeneous the smaller the fragment. Starting from a homogeneous
profile can occasionally result in the emergence of a higher order pattern with a small
basin-of-attraction, see (vi).
Candidates for the expander molecules still have to be identified. However, a first
sequencing screen by our experimental collaborators found several genes which are
expressed in a size-dependent manner. Some of them are even approximately homoge-
neously expressed and show a strong (possibly quadratic) dependence on worm length,
see (iii).
54
3.8 Comparison to experiments in flatworms
Figure 3.5.: The β-catenin concentration forms a gradient from tail to head defining
the AP axis in flatworms. Quantitative Western Blot by Tom Stuckemann in the group
of Jochen Rink (lysate refers to dissolved cell material).
For a homogeneous effect of the expander on the morphogen, the expander might be
a rather small and fast diffusion molecule. Such small molecules could for example
quickly travel between cells via gap junctions. Interesting, it has been reported that the
blocking of gap junctions in flatworms leads to body patterns with a larger wavelength
(152), see (iv).
In summary, we could identify several signatures of our simple model, which help us to
lay out the route for further experiments such as targeted gene knockouts. Thereby, it
is important not to confuse patterns with processes. The same steady state pattern can
arise from very different underlying mechanisms. This is particularly true for the case
of Turing models (147). Already for a simple three component model as we discussed
in this chapter, the knockout of activator or inhibitor might easily provoke misleading
interpretations of the data, see Appendix A.7.4. Considering not an endpoint assay but
a time course of the system after amputations or gene knockouts might to some extent
circumvent this problem. The dynamics typically reveal additional, valuable details
that might enable us to distinguish between different mechanisms.
55
12345678910051015AP axis positionpg β-catenin / μg lysate3. SCALING & REGENERATION OF SELF-ORGANIZED PATTERNS
3.9. Summary and discussion
In this chapter, we introduced a new class of pattern forming mechanisms, which gene-
rate patterns that perfectly scale with system size in a completely self-organized manner.
The patterning scheme relies on expander molecules that dynamically adjust the degra-
dation rates of morphogens in a Turing system. Thereby, the expander controls the
pattern length scales and the source size of the resulting Turing patterns. Conversely,
the expander concentration is dynamic itself and is regulated by the concentrations
of the Turing morphogens. For the feedback introduced here, the relative source size
at steady state is always independent of the system size, see Eq. 3.6. Furthermore,
we showed that a head-tail polarity pattern with a single source region scales as a
function of the system size, is stable with respect to perturbations, and regenerates in
amputation fragments.
Previous works have already discussed scaling of morphogen profiles and patterns in
pre-patterned systems. Either the morphogen source was specified by other pre-existing
cues (24, 26) or the morphogen profile was self-organized according to a Turing mecha-
nism but scaling relied on a predefined geometry (90, 96, 148, 154). Such frameworks
are challenged by the regeneration capabilities of flatworms. In addition, we will show
in Chapter 5 that the length-width ratio of the worms does not stay constant during
growth, which is also not compatible with mechanisms relying on a fixed geometry.
Now, our self-organized model for pattern scaling is able to read out system size in
the absence of any pre-patterns and without strong dependence on specific geometrical
features.
Certainly, body plan patterns in flatworms do not emerge in the total absence of po-
tential pre-patterning cues. For example, it has been shown that the contact of dorsal
and ventral tissue at the wound site causes a regeneration response (102). Still, our
self-organized patterning system captures main characteristics of body plan scaling and
regeneration in flatworms and allows to discuss generic concepts. Thereby, it provides
the basis for further investigations and more sophisticated modeling approaches.
In the minimal theory formulated here, we neglected spontaneous expander degradation.
Such spontaneous degradation would cap the expander concentration and set a lower
size limit for pattern scaling. Additionally, in size-monitoring systems as considered
here, a key challenge relates to the simple fact that these obviously require long-range
communication across the scale of the system. This implies a trade-off between an
upper size limit for scaling, and the time-scale of pattern formation. Here, this time
scale is set by morphogen diffusion and system size. For example, assuming a maximum
56
3.9 Summary and discussion
diffusion coefficient of 100 µm2/s and a maximum organism size of 20 mm, relevant
for the flatworms considered, we infer a patterning time scale of 3−30 days, roughly
consistent with the experimental range of 1−2 weeks for the restoration of body plan
proportions after amputation (4, 140). Note that active, non-directional transport in
addition to passive diffusion could accelerate morphogen dispersal and thus allow for
faster pattern formation (76).
In fact, it has been suggested for flatworms that fast long-range communication might be
implemented via gap junctions or the central nervous system (9, 152, 157). Besides the
transmission of actual nervous signals, a molecular transport along the nerve cords was
proposed (9, 169, 245). Moreover, a gradient of membrane voltage has been discussed
to mediate long range patterning and scaling of body parts (22, 23, 115). Yet, even
when considering faster transport processes than normal diffusion, eventually we can
expect a physical limit to scaling, which would be revealed by the emergence of higher
order body plan patterns. Interestingly, it has been reported that a second head forms
in very large animals of some species (50, 98).
So far, we were mainly concerned with the Wnt/β-catenin system and the gradient for-
mation along the anterior-posterior (AP) axis. Yet, there are also two additional axes,
which are specified by other signaling molecules (141). For example, the Admp/Bmp
system appears to be responsible for the establishment of a dorsal and a ventral side
(61). In flatworms, these molecules seem to cross-react analogous to a Turing system.
Previously, they have also been proposed to be a Turing pair in the frog Xenopus laevis
(122) and a shuttling mechanism has been discussed as a means to adjust the Bmp
gradient to the size of the frog embryo (25).
The theory developed in this chapter provides a framework to understand robust axis
formation and scalable body plan patterning. In the future, it will be important to test
the generic concepts presented here in regeneration experiments and to quantify spatial
profiles of signaling molecules and genetic activity. With this, we aim to identify the
key modules leading to scaling and regeneration.
57
4. Flatworm shape dynamics and motility
4.1. Modes, movement and morphology
While the previous chapters were concerned with the patterning of a specific body
layout, this chapter aims to analyze motility-induced and inter-species variations of the
body shapes as a first step to relate form and function. Ultimately, we would like to
understand why worms look and behave the way they do. Most of the results of this
chapter have been published in Werner et al. (238).
Evolutionary adaption to a specific environment manifests itself in the emergence
of characteristic body morphologies, which among others allow the organism to effi-
ciently sense its environment and steer its path. Thus, motility phenotypes reveal
abnormal body plan patterns. Flatworms usually display a smooth gliding motility,
"Whether it be the sweeping eagle in his flight, or
the open apple-blossom, the toiling work-horse,
the blithe swan, the branching oak, the win-
ding stream at its base, the drifting clouds, over
all the coursing sun, form ever follows function,
and this is the law." -- Louis H. Sullivan,
The Tall Office Building Artistically Considered,
1896 (210)
resulting from a coordinated beating of the cilia
in their densely ciliated ventral epithelium (8, 169,
179, 180). For the cilia to act in synchrony, they
have to be oriented in parallel to each other, which
is coordinated by a structural polarity of the epi-
thelial cells in the plane of the epithelial tissue.
This planar cell polarity system (PCP) comprises
several molecules that asymmetrically localize to
the anterior or posterior side of each cell, respectively, and bind through the membrane
to the corresponding counterpart in the neighbouring cells (51, 196, 226). It has been
described that the PCP system, which is based on local cell-cell interactions, couples to
global polarity cues provided by organism-scale / tissue-scale concentration gradients
of signaling molecules as discussed in the previous chapter (51, 196, 226). Together
with our experimental collaborators, we aim to explore this coupling by quantifying
the gliding speed after knocking out genes of the Wnt/β-catenin pathway as a long
range patterning system.
In a second approach, we analyze and characterize the movement patterns of flatworms
in detail. How do the worms steer their path? We find that a characteristic bending
posture correlates with active turning during gliding, presumably caused by unilateral
58
4.1 Modes, movement and morphology
muscle contractions. Thus, during gliding motility, speed is generated by beating cilia
whereas steering is controlled by muscles. Occasionally, flatworms also show a second
type of motility behavior, called "inch-worming", sometimes also referred to as "peri-
staltic movement" (8, 169, 179, 180). Switching to this motility mode is associated
with escape responses and impaired cilia functionality. We characterize inch-worming
by stereotypic width changes of these worms. Our method reveals regular lateral con-
traction waves with a period of about 4 s in inch-worming worms. Since then, body
contractions have been further analyzed and a recent publication reports on distinct
differences between "scrunching" as a quick escape response and "peristalsis" as a more
persistent gait if cilia functionality is impaired (38). We will compare these results to
our previous findings.
Finally, we employ shape analysis methods to characterize the morphologies of various
flatworm species, which come at very different shapes and sizes. The taxonomic identi-
fication is usually challenging, relying largely on the time-consuming mapping of inter-
nal characters. Here, we illustrate how to systematically and quantitatively distinguish
different flatworm species by head shape. The head is one of the most characteristic
external hallmarks of the flatworm body plan. Many sensory organs for various stimuli
(e.g. light, temperature, touch, chemical signals) are located in the head (21, 95). This
includes two eyespots and the auricles with a high density of nerve cells. Yet, also
other, less prominent shape characteristics might be crucial for the worm to survive,
for example by allowing for more efficient movement and navigation. The systematic
characterization and quantitative comparison of body plan morphologies provides the
basis for relating worm shape to fitness of a species in its specific environmental niche
(103, 199).
We analyzed the motility patterns of Smed and head shapes of different flatworm species
using shape mode analysis based on principal component analysis (PCA) (97, 99, 156).
The technique has been previously applied to biological data to e.g. analyze the dy-
namics of human arm postures, crawling C. elegans worms or swimming sperm cells
(58, 118, 189, 207, 208, 209). In contrast to those examples, which are well described by
their center line, the flatworm body is constantly deforming during movement depend-
ing on the contraction of their muscular plexus due to the absence of skeletal elements
or a rigid body wall. Thus, we adapted PCA to analyze and quantify motility patterns
of a closed non-convex boundary outline.
The starting point of PCA is typically a large data set comprising Na measurements
of Nb features. An example would be the curvature at Nb discrete positions along
59
4. FLATWORM SHAPE DYNAMICS AND MOTILITY
a winding worm at Na time points.
are often correlated, e.g. the curvature at adjacent positions is similar due to a finite
In biological data sets the individual features
stiffness of the body. Therefore, the data might be well described by a much smaller
set of characteristic shape modes. For example a sine wave might be a good first
approximation to the wiggling of a worm. The most descriptive shape modes are
found in a systematic way by determining the largest variations in the data set. At
best, a small number of empirical modes characterizes most of the data and can be
interpreted in a meaningful way. As a side-effect, this method reduces measurement
noise by averaging over several, partially redundant features. PCA can also be seen as
a rotation of the phase-space coordinate system such that it is more suitable for the
data at hand. Afterwards, the data can be projected with only negligible information
loss onto a subset of the phase-space spanned by a small number of axes. These axes
correspond to the shape modes.
4.2. Worm motility reports on patterning defects upon gene
knockout
We analyze movies of gliding worms as described in Appendix C and extract the center
line of the worm body. The midpoint of the center line is tracked to determine the
median worm speed. We observe that wild type worms actively regulate their gliding
speed up to a maximum of 2 mm/s with a weak dependence on size, see black circles
in Fig. 4.1(a). The range in speed is in quantitative agreement with earlier works
in Smed and Dugesia tigrina (155, 212). Next, we compare this result to the speed
of modified worms, for which pathway components of the Wnt/β-catenin patterning
system have been silenced by RNA interference (RNAi). Imaging of wild type worms
has been performed by Nicole Alt under the supervision of the author. The RNAi
experiments have been performed by Sarah Mansour in the group of Jochen Rink (two
RNAi feedings per week). All analysis has been done by the author.
The canonical Wnt/β-catenin signaling cascade is illustrated in Fig. 4.1(b) (82, 107). In
the absence of canonical Wnts, β-catenin is degraded by a destruction complex, which
includes APC as a key component. If Wnt ligands bind to the Frizzled receptors, the
destruction complex is inhibited via Dishevelled (Dvl). Thus, β-catenin can accumulate
in the cell and reach the nucleus to control transcription. Canonical Wnt/β-catenin
signaling has been associated with tail formation (82). Knockout of Dvl by RNAi
feedings leads to the emergence of multiple heads, while knockout of APC leads to the
formation of a second tail after head amputation.
60
4.2 Worm motility reports on patterning defects upon gene knockout
Figure 4.1.: (a) Speed of wild type worms (black circles) and worm speed after APC
RNAi feeding for one week (red dots), two weeks (green dots) and three weeks (blue
dots), respectively (705 wild type worms, 41/38/39 RNAi worms). (b) Canonical Wnt
pathway: binding of Wnt ligands leads to β-catenin translocation to the nucleus (via
Dvl activation and subsequent APC inhibition), adapted with permission from (82).
(c) Quantification of the APC RNAi experiment in panel (a) for small worms (1-3
mm2) and large worms (3-7 mm2). At week 4 worms still show body contractions but
no net displacement anymore. (small: 271 wild type worms, 20/15/14 RNAi worms;
large: 130 wild type worms, 19/22/12 RNAi worms) (d) Speed measurements after 3
weeks of Dvl RNAi feeding (322 wild type worms, 12 RNAi worms of size 1-4 mm2).
(e) Speed measurements after 3 weeks of RNAi of the non-canonical Wnt2 and Wnt5
(154 wild type worms, 17 Wnt2 RNAi worms and 25 Wnt5 RNAi worms of size 2.2-
5 mm2).
(Imaging of wild type worms by Nicole Alt under the supervision of the
author. RNAi experiments by Sarah Mansour. All analyses performed by the author.
Asterisks denote the significance level of the p-value test: * 5% and ** 1%. Error bars
correspond to standard deviations.)
61
0510152000.511.52Speed in mm/sMean size in mm2WTW1W2W3W400.511.5Mean speed in mm/ssmallWTW1W2W3W400.511.5Mean speed in mm/sWTWnt2Wnt500.511.52Mean speed in mm/sWTDvl00.511.5Mean speed in mm/sWntFrizzledDishevelled (Dvl)Destruction complex with APCβ-cateninDNA(a)(e)(d)(c)(b)large**************4. FLATWORM SHAPE DYNAMICS AND MOTILITY
In intact worms, we observe a motility phenotype for APC RNAi, see colored points in
Fig. 4.1(a). Worm speed is reduced over the course of the experiment. We can quantify
this trend for two size classes and see a significant effect already after two weeks of
RNAi feeding, see Fig. 4.1(c). It remains to determine whether APC acts directly on
the PCP pathway or whether a graded β-catenin signal is required to maintain planar
cell polarity.
Similarly, when knocking down both Dvl genes of Smed , we observe a reduced speed
suggesting a disoriented cilia carpet, see Fig. 4.1(d). In fact, it is known that Dvl is not
only a component of the canonical Wnt signaling cascade, leading to a translocation
of β-catenin to the nucleus, Dvl also acts in non-canonical Wnt signaling, where it
activates the PCP pathway (8, 60, 107). It has been shown that silencing Dvl leads to
a less, shorter and more disorganized cilia (8, 9).
In order to determine to what extent non-canonical Wnt signaling is related to PCP in
flatworms, Sarah Mansour performed RNAi of Wnt2 and Wnt5. Unlike the canonical
tail Wnts, Wnt2 is expressed in the head and Wnt5 is expressed around the body
margin and especially along the central nervous system (9, 81). Especially Wnt5 has
been discussed to act in non-canonical signaling in flatworms and is also involved in
PCP systems of other organisms (9, 59). Fig. 4.1(e) shows no movement phenotype for
Wnt2 RNAi, but Wnt5 RNAi results in a significant reduction in speed. Interestingly,
the Wnt5 phenotype is rather subtle. While for APC and Dvl RNAi, the worms seem
to loose the ability to glide, switch to inch-worming motility and eventually cease being
motile altogether, the Wnt5 RNAi worms are still capable of normal gliding motion, just
slower. It is an interesting question, whether this is due to an incomplete knockdown
or the particular effect of Wnt5 on the PCP system.
In summary, analyzing worm movement upon RNAi is a simple and non-lethal approach
to identify patterning phenotypes. The method reveals even subtle effects as for Wnt5
and inspires further more elaborate measurements of cilia orientation and gene expres-
sion profiles.
62
4.3 Shape mode analysis of 2D worm outlines
4.3. Shape mode analysis of 2D worm outlines
Next, we analyze the behavioral repertoire of flatworms in greater detail. Different
movement strategies are reflected in bending of the worm body. Due to the large body
plasticity, motility-induced changes in body posture are not well described by only
the center line but require an analysis of the perimeter dynamics. Thus, we face the
challenge of characterizing the shape of closed, planar curves. For non-convex shapes,
this can be non-trivial.
We extract the midline and the perimeter of the worms from movie sequences as de-
scribed in Appendix C, see Fig. 4.2(a). The imaging has been performed by Nicole Alt
under the supervision of the author. The worm shape corresponds to a closed curve
described by a position vector r(s) as a function of arc-length s along its circumference,
see Fig. 4.2(b). We use the tip of the worm tail as a distinguished reference point r1 that
specifies the position of s = 0. We further specify a center point r0, using the midpoint
of the tracked center line of the worms. The profile of radial distances ρ(s) = r(s)− r0
measured with respect to the center point r0 characterizes outline shape, even for non-
convex outlines. Shapes of convex curves might also be characterized by a profile of
radial distances ρ(ϕ) as a function of a polar angle ϕ. However, this definition does
not generalize to non-convex curves (or, more precisely, to curves that are not radially
convex with respect to r0). To adjust for different worm sizes, we normalize the radial
distance profiles by the mean radius ρ = (cid:104)ρ(s)(cid:105) as ρ = ρ(s)/ρ and plot it as a function
of normalized arc-length s = s/L, where L is the total length of the circumference.
As a mathematical side-note, we remark that using the signed curvature κ(s) = (∂2
s r(s))·
(∂sr(s)) along the circumference, instead of the radial distance profile ρ(s), would
amount to a significant disadvantage: The property that a certain curvature profile
corresponds to a closed curve imposes a non-trivial constraint on the set of admissi-
ble curvature profiles. For the normalized radial distance profiles, however, there is
a continuous range of distance profiles that correspond to closed curves, making this
choice of definition more suitable for applying linear decomposition techniques such
as shape mode analysis. In fact, given a particular normalized radial distance profile,
the corresponding circumference length L/ρ is reconstructed self-consistently by the
requirement that the associated curve must close on itself, see Appendix D.1.
63
4. FLATWORM SHAPE DYNAMICS AND MOTILITY
Figure 4.2.: Three shape modes characterize projected flatworm body shape dyna-
mics. (a) Our custom-made MATLAB software tracks worms in movies and extracts
worm boundary outline (red) and centerline (blue). (b) The radial distance ρ(s) be-
tween the boundary points and midpoint of the centerline (r0, red dot) is calculated
as a parameterization of worm shape. We normalize the radial distance profile of each
worm by the mean radius ρ. (c) Covariance matrix of the radial distance profiles: the
second symmetry axis (dotted line) corresponds to statistically symmetric behavior of
the worm with respect to its midline (745 worm movies, imaging by Nicole Alt under
the supervision of the author, all analysis performed by the author). (d) The three
shape modes with the largest eigenvalues account for 94% of the shape variations. The
first shape mode characterizes bending of the worm and alone accounts for 61% of the
observed shape variance. We show its normalized radial profile on the left as well as
the boundary outline corresponding to the superposition of the mean worm shape and
this first shape mode (solid red: B1 = 1, dashed red: B1 = −1, black: mean shape
with B1 = 0). The second shape mode describes lateral thinning (B2 = 0.3), while
the third shape mode corresponds to unlike deformations of head and tail (B3 = 0.8),
giving the worm a wedge-shaped appearance. (e) The first shape mode with score B1,
describing worm bending, strongly correlates with the instantaneous turning rate of
worm midpoint trajectories. (f) We manually selected 30 movies where worms clearly
show inch-worming and 50 movies with no inch-worming behavior. The variance of
score B2 and B3 increases for the inch-worming worms. (g) The autocorrelation of
mode B3 and the crosscorrelation between mode B2 and mode B3 reveals an inch-
worming frequency of approximately 1/4 Hz, hinting at generic behavioral patterns.
64
−2−1012−0.4−0.200.20.4Turning rate (1/s)Score B1 of first moder0r1 tailheadtail −0.2−0.100.10.2Third shape mode v3 tailheadtail −0.2−0.100.10.2Second shape mode v2 tailheadtail −0.2−0.100.10.2First shape mode v160.7 %24.8 %8.8 %(a)(d)(b)correlation 0.52(e)(c)1 mm(f)tailtailhead(g)0246810−0.5−0.2500.250.50.751correlationtime interval (s)autocorrelation B3crosscorrelation B2 - B3no inchworminginchworming00.10.20.30.4variancevar(B2)var(B3)var(B2)var(B3)stailtailhead4.4 Shape dynamics during crawling and inchworming
4.4. Shape dynamics during crawling and inchworming
4.4.1. A bending mode and two width-changing modes
We extracted Nw = 29 993 worm outlines from a total of 745 analyzed movies. We
computed normalized radial distance profiles as described above, each profile being
represented by Nr = 200 radii, resulting in a large Nw × Nr data matrix Ri,j = ρi(sj).
From the average of all radial profiles, we define a mean worm shape that averages out
shape variations ρ0(s) =(cid:80)Nw
we can devise a Nw × Nr matrix R0
the mean profile. Next, we computed the Nr × Nr covariance matrix
i=1 ρi(s)/Nw, see Fig. 4.2(d) (right inset, black). From this,
i,j = [ρ0; ...; ρ0], for which all the rows are equal to
C = (R − R0)T (R − R0)
(4.1)
between the individual radial profiles, using the centered (mean-corrected) data ma-
trix, see Fig. 4.2(c). The covariance matrix is by construction symmetric along the
dashed line. The approximate symmetry of the covariance matrix along the dotted
diagonal shows that shape variations are statistically symmetric with respect to the
worm midline. For example the worm bends as often to the left as to the right.
The Nr eigenvectors vj(s) or shape modes of this covariance matrix correspond to
axes of a new coordinate system. In this coordinate system, the variation of the data
along each axis is linearly uncorrelated and the corresponding variance is given by the
respective eigenvalue. The shape modes with the largest eigenvalues are those with
maximal descriptive power. We can uniquely express the worm shape as
Nr(cid:88)
ρ(s) =
Bjvj(s) ,
(4.2)
j=1
where the shape scores Bj can be computed by a linear least-square fit. Due to corre-
lations, many data sets can be well described by a truncated sum, using only a small
number of shape modes with the largest eigenvalues.
Fig. 4.2(d) shows the first three shape modes, which together account for 94% of the ob-
served variation in shape. We find that the dominant shape mode v1 is anti-symmetric,
describing an overall bending of the worm. In contrast, the second and third mode de-
scribe symmetric width changes of the worm: The second shape mode v2 characterizes
a lateral thinning of the worm associated with a pointy head and tail. Correspondingly,
a negative contribution of the second shape mode with B2 < 0 describes lateral thicke-
ning of the worm (with a slightly more roundish head and tail). The third shape mode
65
4. FLATWORM SHAPE DYNAMICS AND MOTILITY
v3, finally, is associated with unlike deformations of head and tail, giving the worm
a wedge-like appearance. Superpositions of these three shape modes describe in-plane
bending of the worms, and a complex width dynamics of head and tail.
We find that shape changes control the direction of gliding motility and thus steer the
worm's path: Fig. 4.2(e) displays a significant correlation between the rate of turning
along the worm trajectory and the first shape score B1, which characterizes bending
of the worm. The sign and magnitude of this "bending score" directly relates to the
direction and rate of turning. For simplicity, we restricted the analysis to a medium
size range of 8-10 mm length, analogous results are found for other size classes.
4.4.2. The second and third modes characterize inch-worming
In addition to cilia-driven gliding motility, flatworms employ a second, cilia-independent
motility pattern known as inch-worming, which provides a back-up motility system in
case of dysfunctional cilia (8, 169, 179, 180). We test whether modes two and three
might relate to this second motility pattern. For this, we analyzed movies of small
worms known to engage more frequently in this kind of behaviour. We manually clas-
sified 80 movies of worms smaller than 0.9 mm that had been starved for 10 weeks,
yielding a number 30 inch-worming and 50 non-inch-worming worms for a differentiated
motility analysis (cases of ambiguity were not included). We find that the second and
third shape mode, which characterize dynamic variations in body width, are indeed
more pronounced in inch-worming worms, see Fig. 4.2(f). Next, we computed the tem-
poral autocorrelation of time series of the second shape mode B3, see Fig. 4.2(g) (solid
blue). We observe stereotypical shape oscillations with a characteristic frequency of
0.26 Hz. From the cross-correlation between B3 and B2 in Fig. 4.2G (dashed black),
we find that both shape scores oscillate with a common frequency and relative phase lag
of −0.6 π (where B2 lags behind). Thus, both shape modes act together in an orches-
trated manner to facilitate inch-worming, hinting at coordinated muscle movements
and periodic neuronal activity patterns.
4.5. Discriminating flatworm species by shape
After having developed tools to measure shape changes of the same species over time, we
next explored the utility of shape mode analysis for comparing different species. The
model species Schmidtea mediterranea is but one of many hundred flatworm species
existing worldwide (50, 116). We aim to introduce shape mode analysis as a technique
66
4.5 Discriminating flatworm species by shape
Figure 4.3.: (a) Application of our method to parametrize head morphology of four
different flatworm species. For each species, time-lapse sequences of 4 different worms
were recorded as two independent runs of 16 frames each. The head is defined as
the most anterior 20% of the worm body. Radial distances ρ(s) are computed with
respect to the midpoint of the head (red dot at 10% of the worm length from the
tip of the head). (b) By applying PCA to this multi-species data set, we obtain two
shape modes, which together account for 88% of the shape variability. Deformations
of the mean shape with respect to the the two modes are shown (black: mean shape,
red: superposition of mean shape and first mode with B1 = ±0.4 and second mode
with B2 = ±0.2, respectively). We represent head morphology of the four species in
a combined shape space of these two modes. Average head shapes for each species
are indicated by crosses, with ellipses of variance including 68% (dark color) and 95%
(light color) of motility-associated shape variability, respectively. (Imaging by Nicole
Alt and Miquel Vila-Farr´e, data analysis by the author)
to facilitate taxonomic classification of these species. PCA enables us to extract typi-
cal shape variations between species without "a priori" assumptions on characteristic
features.
Having available a large live collection of flatworm species, we choose four species repre-
senting the genera Girardia, Phagocata, Schmidtea and Polycelis. Besides potentially
size-dependent variations in aspect ratio, the four species differ by their characteristic
head shapes, see Fig. 4.3(a). Accordingly, we restrict the shape analysis to the head
region only (defined as the most anterior 20% of the worm body). In analogy to the
procedure described above for the full worm body, we characterize each head shape by
a vector of distances from the midpoint of the head (red dot, 10% of the worm length
from the tip of the head) to the outline ρ(s) of the head region. Next, PCA is applied
to the normalized radial distance profiles like before.
67
GirardiaPhagocataSchmidteaPolycelis1 mm−0.500.5−0.200.2Score of second mode B2Score of first mode B1(a)(b)4. FLATWORM SHAPE DYNAMICS AND MOTILITY
We found that the first two eigenmodes captured 88% of head shape variability within
this multi-species data set. Fig. 4.3(b) shows species-specific mean shapes for each
of the four species in a combined head shape space, as well as ellipses of variance
covering 68% (dark color) and 95% (light color) of motility-associated shape variability,
respectively. Shape reconstruction from individual modes is explained in Appendix D.2.
The first mode seems to describe the extent of the auricles while the second mode can
be associated with the auricle position. This comparison of flatworm species represen-
ting four genera illustrates linear dimensionality reduction as a simple means to map
morphological differences across species.
4.6. Discussion
In this chapter, we have discussed several approaches to relate form and function in
flatworms focussing on motility. Movement patterns are generated by various features
of the body plan such as cilia and muscle layout. By analyzing worm motility, we gained
first insights into their functionality and coordination. In a first approach, we measured
worm speed in wild type and RNAi treated animals and could quantify subtle motility
phenotypes, which hint at patterning defects of the PCP system.
In a second approach, we analyzed the shape variations during movement, adapting
principal component analysis in order to apply it to 2D outlines. We characterized
inchworming as a second motility mode, different from cilia-based gliding motility,
which is driven by well-coordinated muscle contractions with a characteristic frequency
of about 0.26 Hz. In a recent work, contractile motility gaits have been further analyzed
(38). Cochet-Escartin et al. distinguish between two behavioral traits: (i) scrunching
as a transient and fast response to cutting as well as electrical, acidic and temperature
shocks, (ii) peristalsis as a more persistent and slow movement strategy if cilia func-
tionality is impaired. The authors could induce the latter in two different ways: (i)
by RNAi of the iguana gene leading to defective ciliogenesis and (ii) by increasing the
viscosity of the media (by adding 16% ficoll).
Interestingly, they extracted frequen-
cies of the area oscillations during peristalsis and obtain 0.28 ± 0.03 Hz (iguana) and
0.26 ± 0.02Hz (ficoll), respectively. These values are very similar to our result. To-
gether, the three conditions illustrate a very generic inchworming/peristalsis behavior
despite their higher level of complexity compared to other model organisms such as C.
elegans (173, 184, 208, 244). It remains to determine how this relates to the underlying
structure of the muscular plexus and the nervous system.
68
4.6 Discussion
In a third approach, we apply PCA to head shapes of different flatworm species as a tool
for taxonomic classification. We captured the main characteristics of head shapes of four
species belonging to different genera by just two shape modes. The result suggests that
head morphogenesis is mainly controlled by two molecular networks: one to determine
the head width and a second one for the anterior positioning of the auricles. The
availability of transcriptome sequence data for these species will now provide us with
the opportunity to test this hypothesis. We also have planned to extend the analysis
to many more species of the collection in the laboratory of Jochen Rink and relate
the characteristic body plan features to the conditions in the respective environmental
niche. A similar inter-species comparison with respect to evolutionary selected traits
has been performed for beaks of Darwin finches, phalanxes of vertebrates, heads of ants
and wings of bats (103, 199). Based on these analyses, we expect for flatworms that a
small number of shape modes with the largest eigenvalues spans the feature space, in
which most of the shape variations take place. Highly specialized species have typically
rather extreme morphologies and can be found at the corners of the observed shape
set.
In contrast, generalists show mixtures of different traits, which simultaneously
optimizes the fitness to fulfil several tasks. It will be interesting to test this hypothesis
for flatworm species, some of which inhabit extreme environments.
69
5. Quantitative study of
flatworm growth and cell turnover
5.1. Homeostasis is a dynamic steady state
Multicellular organisms in their adult stages have a relatively constant outward ap-
pearance but the integrity of a functional body is only maintained due to a permanent
replacement of damaged or lost cells (158). In humans, it has been estimated that the
total mass of cells we replace every year is almost as much as our entire body weight
(166). The time scales of the cellular turnover range from a few days for blood cells
and the stomach to several years for cells in the heart and the skeleton (124). Only a
small fraction of cells like some nerve cells might never be replaced.
From the point of view of dynamical systems theory, a constant outward appearance in
the face of permanent turnover can be considered as a steady state, for which inflows
(i.e. the generation of new cells) and outflows (i.e. the loss of old cells) are balanced.
"ποταοσι τοσιν ατοσιν βαίνουσιν
τερα κα τερα δατα πιρρε." (On those
stepping into rivers staying the same other and
other waters flow.) -- Heraclitus of Ephesus,
500 BC (105)
The underlying processes of cell division and cell
loss have to be well controlled in order to avoid a
disintegration of the organism or an uncontrolled
overgrowth such as cancer. For this to function
robustly, there is likely a communication between
the dying cells in the tissue and the dividing stem cells in their niche, see Fig. 5.1. A
limited capacity to replace cells is related to sickness and aging of organisms (158).
During growth, the balance between inflows and outflows is shifted to an increased cell
division in comparison to cell loss.
Importantly, growth is not only controlled by a
developmental program but also influenced by the availability of food (113, 244). In
many organisms, nutrition levels affect the speed of growth and the final size.
In this chapter, we address the link between the microscopic scale of cell turnover and
the macroscopic growth in our favourite multicellular organism the flatworm Smed .
Previous research has mainly been devoted to cell death and cell division as individual
processes but little is known about how these two processes are jointly regulated and
mutually influence each other to result in coordinated growth at the organism scale
70
5.1 Homeostasis is a dynamic steady state
Figure 5.1.: (a) Stem cell division and the loss of cells have to be balanced to maintain
a constant size of the tissue.
(b) For growth and degrowth, the balance of these
processes is slightly shifted in a well controlled manner.
(158). Flatworms are suitable model organisms to study cell turnover and growth in
a comprehensive way as they permanently replace all of their cells within a few weeks
and reversibly grow over a 40-fold range in size (16, 140, 151, 158, 170). Thereby,
the cell size appears to be approximately constant and the reversible growth mainly
corresponds to a change in cell number (16, 140, 150, 193). Interestingly, changes in
worm size has been previously related to aging and small worms have been described
to be more juvenile. In Appendix E.1, we discuss signatures of aging in sexual and
asexual flatworms.
Two basic models of growth and turnover control have been discussed for flatworms:
(i) stem cell control and (ii) cell death control (159), see Fig. 5.2. In the first scenario,
dividing stem cells or their descendants induce cell death in differentiated cells for
replacement. In order to generate growth, the replacement must not be perfect. We
might hypothesize that cells only respond with a certain probability (maybe depending
on their age or fitness) to the death signal from the stem cell pool.
In the second
scenario, the dying cells send out signals to enhance cell division and recruit progenitor
cells to the tissue. In both cases, an additional input from the nutritional status of the
worm is needed for growth.
By studying flatworms as a model system, we aim to bridge the scales between cellular
turnover processes and growth and degrowth on the organismal level. First, we quantify
71
(a) Replacement & Repair?divisionlossstem cellsdifferentiated tissue(b) Growth & Degrowthlossdivision5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.2.: Two paradigms of turnover control: (a) stem cells and their descendants
induce the death of differentiated cells, (b) dying cells induce cell division and recruit
progenitor cells to the tissue. In both cases feeding enhances cell division.
the size of the worms as well as growth dynamics and find a non-trivial size-dependence
of the growth and degrowth rates. All experiments were performed in the laboratory
of Jochen Rink. Several members have been involved in this close collaboration as we
will explicitly state for each data set. In a second step, we theoretically discuss three
paradigmatic models for how cell division and cell loss might be linked to the metabolic
status of the worm. Each model is able to explain the observed behavior within the
accuracy of the measurements. We obtain several predictions of the models that can be
tested in further experiments. The results directly relate to the questions of how these
worms can survive long starvation periods and what determines the limits to growth.
Finally, we establish the theoretical framework for future measurements to reveal the
control logic of turnover processes.
5.2. Size-dependent growth and degrowth dynamics in
flatworms
5.2.1. Allometric scaling laws
5.2.1.1. Measurements of area, cell number and mass of the worm
The first step towards measuring growth dynamics is to accurately determine worm size.
Together with our experimental collaborators we performed experiments to extract the
outer dimensions like length and width as well as the cell number and the mass of the
worms. From these measurements we obtained allometric scaling laws that relate the
various quantities for cross-validation and to gain further information about the body
plan of Smed, see Fig. 5.3.
One important quantity is the area of the worm, which refers to the 2D-projection
of the worm body. Area as well as other outer dimensions are readily accessible by
microscopic imaging and we have developed a protocol to measure these quantities in
72
(a) stem cell control(b) cell death control5.2 Size-dependent growth and degrowth dynamics in flatworms
a precise and reproducible way, see Appendix C. Several student helpers in the group
of Jochen Rink have been involved in the imaging of the worms. The author designed
the experiments, provided training and supervision and performed the analysis of the
data.
Fig. 5.3(a) reveals an allometric scaling relation between worm area and length with
an exponent of 1.81± 0.01. Thus, large worms are relatively thinner than small worms.
Remarkably, the scaling law holds across the entire size range with a very small trans-
verse spread and a high reproducibility. These worms were starved for at least a week.
The respective scaling law for well fed worms can be found in Appendix E.2. There we
also discuss the details of the data analysis. The scaling exponents have been deter-
mined using a robust regression algorithm with bi-squared weights implicitly obtained
from the spread of the data.
Ultimately, we are most interested in the changes in cell number because this quantity
directly relates to the rates of cell division and cell death. However, measurements of cell
numbers are fatal for the worms. As a solution, we establish a functional relationship
between cell number and worm area. By using the worm area as a read-out for size,
we can reduce the number of worms needed for various experiments and perform time-
course experiments with well-defined initial sizes.
The cell number itself has been measured in two different ways by Albert Thommen in
the group of Jochen Rink. First, he determined the amount of histones in the worm.
Histones are structural units that organize the DNA in eukaryotic cells (68). The DNA
strands are wrapped around the histones in the nuclei for compaction. Histones are
only synthesized during cell division and the amount of histones was measured to be
constant throughout the lifetime of the cells. After the amount of histones per cell has
been determined, the total amount of histones in a worm can be used as a read-out for
cell number. For this, the amount of histones was measured via Western blotting in a
known fraction of the total protein mass. In a second experiment, cell numbers were
obtained by disintegrating the worm into individual cells and automatically counting
cells in microscope images of a well defined volume fraction of the worm.
The resulting cell numbers are plotted in Fig. 5.3(b) as a function of worm size. We
find that the number of cells scales almost linearly with the area of the worms with a
scaling exponent of 1.11± 0.02. The most simple explanation would be that the height
hardly changes and also the large worms remain rather flat. Alternatively, the size of
cavities like the gut might increase over-proportionately, compensating for an potential
increase in height for larger worms. Furthermore, it might be possible that the fraction
73
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.3.: Allometric scaling laws: (a) Width and length of the worm do not change
proportionately. Large worms are thinner than small worms (Imaging by Nicole Alt
under the supervision of the author, analysis by the author, 722 measurements of star-
ving worms). (b) Cell numbers are measured by direct cell counting (green circles)
and by histone quantifications (blue dots). The cell number changes almost propor-
tionally with worm area (Imaging and cell number quantifications by Albert Thommen,
analysis by the author, 31 and 45 worms, respectively). (c) In contrast, the dry mass
increases stronger than area and cell number (Imaging and mass measurements by
Albert Thommen, analysis by the author, 21 worms).
74
051005101520Area in mm2Length in mmLog lengthLog area(a)01020304002468Area in mm2Cell number in 106(b)05101502004006008001000(c)Area in mm2Dry mass in µgLog cell numberLog areaLog areaLog dry mass5.2 Size-dependent growth and degrowth dynamics in flatworms
of cell types changes, which might also explain a linear scaling despite variations in
height. Irrespective of which scenario applies, the scaling exponent shows that our area
measurements are a good proxy for cell number.
Finally, we also determine the dry mass of the worms. Again, the experiment has been
carried out by Albert Thommen, while the analysis of the movies has been performed
by the author. Although more data is needed for a more reliable result, the analysis of
the first data points suggests that the scaling exponent of 1.33 ± 0.06 for the dry mass
is larger than the exponent for the cell number. Therefore, the mass per cell seems to
slightly increase with size. Interestingly, some preliminary measurements suggest that
in contrast the protein mass per cell is independent of worm size, see Appendix E.2.
5.2.1.2. Comparison and interpretation of scaling laws
Previously, contradictory results have been published on scaling laws in Smed and other
related species, see Tab. 5.1. For Planaria maculata, the scaling of worm mass with
area to the power of 1.38 agrees well with our result (227). For Dugesia lugubris, Lange
measured the volume in serial sections and found a quadratic relationship with length,
which is in agreement with our scaling law for the cell number (110, 111). Bagun`a
et al. have measured cell number and worm volume in Smed (13, 16). In agreement
with our data, they find similar values for the cell numbers and from them we can
obtain a quadratic scaling with the length of the worm. However, a few years later,
the same group published various scaling laws for four different species including sexual
and asexual flatworms which do not agree with our results.
Further, more recent measurements of the worm mass by Oviedo et al. also deviate
from our scaling law (150). However, a scaling with L0.9 is a rather questionable result
considering that the worm width also increases. In the same paper, the authors counted
the number of cintillo cells, which are found around the margin of the head and are
involved in mechanosensing. If we assume that the head can be approximated by a
semicircle and that the number of cintillo cells increases proportionately to the head
margin, the width of the worm at the head follows L0.68, see Appendix E.2, which
is similar to our results. For Schmidtea polychroa, the wet weight and the protein
content have been measured (133). At least the protein content might show a similar
dependence on area as in our experiment. Finally, Takeda et al. measured the total
DNA mass in Dugesia japonica and find an approximately linear dependence on body
length (211). As the cell number can be assumed to be roughly proportional to the total
DNA mass, this is not in agreement with our measurements in Fig. 5.3(b). In analogy
75
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
scaling law
M ∝ A1.38
V ∝ L2.02
N ∝ L2
N ∝ L1.81
N ∝ A1.44
A ∝ L1.26
Mwet? ∝ L0.9
Mwet ∝ A
Mprot ∝ A
N ∝ L
species
our result (Smed )
reference
Planaria maculata ?
Dugesia lugubris†
Smed
Smed
Smed
Schmidtea polychroa†
Schmidtea polychroa†
Dugesia japonica†
M ∝ A1.33
(N ∝ L2.01)
N ∝ L2.01
N ∝ L2.01
N ∝ A1.11
A ∝ L1.81
M ∝ L2.40
M ∝ A1.33
Mprot ∝ A1.13
N ∝ L2.01
(227)
(110, 111)
(13, 16)
(177)
(150)
(133)
(133), Appendix E.2
(211)
Table 5.1.: Scaling laws for various flatworm species found in the literature. Our
experiments are done in asexual worms, † denotes the sexual strains, ? marks a species,
which cannot be taxonomically classified.
to the argument about the worm mass of Oviedo et al., such a linear relationship of
cell number with length is unlikely given the significant increase in worm width in
larger worms. Note that the spectroscopic quantification of flatworm DNA is rather
unreliable because it interferes with the absorption peak of the pigments. Hence, we
apply different approaches.
Thus, while we could confirm several measurements of scaling laws in flatworms ob-
tained before 1980, we find much less agreement with recently published results. This
discrepancy with respect to our data might only in parts be explained by species-
specific variations and differences between sexual and asexual strains and could relate
to the general problem of size measurements in animals with flexible body shapes. In
our experiments, we have paid particular attention to accuracy and reproducibility by
analyzing movie sequences of individual worms, following a strict protocol. Exten-
sive cross-checks as well as consistency-checks between redundant measurements were
performed.
Our results on the scaling of the cell number suggest that flatworms indeed stay rather
flat. As the worms lack a blood system, this would ensure that all cells can be provided
with oxygen by diffusion from the outer epithelium. While this is the most likely
yet not the only explanation for the observed scaling relation in Fig. 5.3(b), further
measurements of worm heights are necessary to confirm our hypothesis. More data
points are also needed for the scaling relation of the dry mass. The data at hand
76
5.2 Size-dependent growth and degrowth dynamics in flatworms
suggests that the characteristics of the cell population depend on size in the sense that
in larger worms there are more cells with an increased mass. Note, however, that a first
approximate estimation of protein content in the worm does indicate that the protein
content per cell roughly stays constant, see Appendix E.2. Thus, a possible explanation
for the increase in cell mass could be the storage of lipids as energy resources in larger
worms.
Having established size measurements, next, we aim to analyze growth dynamics.
5.2.2. Characterizing the immediate growth response upon feeding
Worms respond to feeding with a fast mitotic peak within the first 12 hours that lasts
for 3 − 5 days (12, 13, 16, 71, 139, 142). We would like to relate this mitotic feeding
response to the increase in worm size. Fig. 5.4(a) shows the changes in worm area
immediately after feeding. After an initial increase of the worm area, the worms shrink
again and return to their initial sizes within approximately two weeks.
Note that the worm area already approaches its maximum very rapidly within the
first 12 hours after feeding, although the mitotic response lasts for following next days.
Thus, we were wondering which part of the peak can be explained by pure stuffing
of the gut and which part reflects the actual cell division response. To this end, we
measured the growth response also for irradiated animals without stem cells, see red
curves in Fig. 5.4(b). Interestingly, the stuffing effect last for at least the first 4 days
until the irradiated worms dissociate. Thus, the initial increase in worm area after
feeding actually reflects stuffing.
Nevertheless, these measurements provide us with the following information: First,
the food fills the gut for at least 4 days, most likely providing the dividing cells with
nutrients during this time. Second, the growth response appears to be rather generic
and food intake is approximately proportional to worm size. We also have not found an
obvious dependence on feeding history. Growth peaks after longer starvation periods
look similar to the results above, see Appendix E.2. Finally, the growth effect decays
within two weeks. Thus, worms should approximately keep their size when being fed
every two weeks. After one week, the worm size has been roughly increased by 2%.
Further experiments will aim to extract changes in cell number upon feeding to reveal
the actual growth response.
77
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.4.: Growth response upon feeding (after one week of starvation): (a) The
worm area increases rapidly within the first day and returns again to its initial value
after 2 weeks. The area is normalized by the average of day 0−7. The individual worm
tracks are shown in gray, the black line illustrates the average behavior across all worms
(Imaging by Johanna Richter, Jordan Ferria and Nicole Alt under the supervision of
the author, analysis by the author, 21 worms). (b) Irradiated worms (red) and non-
irradiated worms show the same initial increase in area due to intake of food. The
area is normalized by the average of day 0− 4 (Imaging by Ashutosh Mishra under the
supervision of the author, analysis by the author, 5 irradiated and 5 non-irradiated
worms each).
78
0h12h24h36h2d3d4d5d0.70.80.911.11.2Normalized areaTime after feeding1.3 - 2.3 mm20h12h24h36h2d3d4d5d0.70.80.911.11.2Time after feeding6 - 9 mm2−10123456789140.70.80.911.11.2Time after feeding (days)Normalized area3 - 6.5 mm2(a)(b)wildtypeirradiated5.2 Size-dependent growth and degrowth dynamics in flatworms
5.2.3. Small worms grow and degrow faster than large worms
5.2.3.1. Measurement of growth rates
Even though we have learned that stuffing effects compromise tracking of the short
term growth pulse, we can still quantify the averaged growth behavior on the time
scale of weeks and in particular analyze how the growth and degrowth rates depend
on feeding. Fig. 5.5(a)-(c) shows the growth tracks for individual worms of different
sizes for three feeding conditions: starvation, feeding every second week and feeding
every week.
In agreement with Fig. 5.4, the worms that are fed every second week
approximately maintain a constant size. One can clearly recognize the growth response
upon feeding that decays until the next feeding event. The concatenation of short-term
growth pulses results in a zig-zag line. For the other two cases, we compute average
growth rates, see Fig. 5.5(d). The data is very noisy for the well-fed worms because
of the stuffing effect and the excretion of digested food. Still, we can obtain a clear
linear trend using two independent fitting procedures. Besides the robust regression of
the growth rates (black, with 95% confidence intervals), we construct a master curve
from the growth tracks assuming negligible effects of the feeding history, see Fig. 5.5(e).
The growth tracks of Fig. 5.5(c) are shifted horizontally in an iterative procedure to
minimize the variance from the average curve. Finally, we fit a Fermi function as the
master curve, which corresponds to a growth rate that linearly depends on size. Both
approaches agree well with each other.
Fig. 5.5(d) shows a clear size-dependence of the growth and degrowth rates. Small
worms appear to grow faster than large worms. Note that even the largest worms have
a positive growth rate, meaning worms usually undergo fissioning before reaching the
limits to growth. Surprisingly, small worms also degrow faster. This trend appears to
be rather independent from the feeding history, see Fig. 5.5(f).
5.2.3.2. Discussion of growth and degrowth dynamics
As for the allometric scaling laws, our results on the growth dynamics rather agree with
very early measurements and less so with more recent publications.
At the beginning of the 20th century, Abeloos investigated growth dynamics in another
flatworm species (1). In analogy to our conclusion, he found that the increase in dry
mass is stronger for smaller species.
Bagun`a et al. have analyzed growth and degrowth dynamics of Girardia tigrina for
various feeding conditions (15, 18). We can extract from these data sets that their
growth rates are also decreasing with size during feeding, see Appendix E.2. The
79
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.5.: (a)-(c) Measurements of worm area for different feeding conditions (Ima-
ging by Nicole Alt und Johanna Richter under the supervision of the author, analysis
by the author includes 66, 41 and 48 worms, respectively). (d) Degrowth rates (red
circles) have been computed from data in panel (a) by linear fits across three windows
with 6 data points each using a robust algorithm. The size dependence is fitted by the
function c1 + c2 A/(c3 + A) with c1 = −3.17 ± 0.38 %/day, c2 = 2.72 ± 0.33 %/day,
c3 = 2.11 ± 1.26 mm2 (black curve). (continued on next page)
80
FMMMMMMM2d17d7d2d17d......FMFMFFMF7d3d3d3d4d4d4d...FMFMFMFM7d7d7d3d3d4d4d(a) Starvation(b) Feeding every second week(c) Feeding every week1471005101520WeeksArea in mm213579010203040WeeksArea in mm22468024681012WeeksArea in mm21471001234WeeksArea in mm2(d)0102030−6−4−20246Area in mm2Area growth rate in percent / day(e) Feeding every weekStarvation(f) 0246−3−2−10Area in mm2Area degrowth rate in %/ dayearly (1./4. week)middle (4./7. week)late (7./10. week)050100150010203040Area in mm2Time in days05010015010−1100101102Time in daysArea in mm25.2 Size-dependent growth and degrowth dynamics in flatworms
Figure 5.5.(continued from previous page): Growth rates (blue cicles) have been
determined from the growth tracks in (c) by an exponential fit across two windows with
5 data points each. The trends in the growth rates are fitted by the robust regression
c1 − c2 A with c1 = 4.18 ± 0.35 %/day, c2 = 0.10 ± 0.03 %/day/mm2 (black line with
95% confidence intervals in gray). A direct fit of the master curve in (e) agrees well
(red).
(e) Growth tracks in (c) can be shifted horizontally (blue) to collapse onto
a master curve (red), which is fitted by the Fermi function c1/(c2 + c3 e−c1t) with
c1 = 4.4 ± 0.3 %/day, c2 = 0.077± 0.012 %/day/mm2, c3 = 19± 5 %/day/mm2. Inset
with log-linear plot illustrates the deviation from an exponential law. (f) Magnification
of the degrowth rates for time intervals after the initial feeding shows no obvious
dependence on feeding history. Error bars represent the standard error of the mean.
growth rates are smaller than in our case, which could be a species-specific effect or
due to gut stuffing. Furthermore, it might also be related to the lower temperature of
12◦C, while our experiments are conducted at 20◦C. More importantly, the degrowth
rates during starvation seem not to depend on the sizes of the worms in contrast to our
data.
Other groups have also measured growth dynamics and typically fitted exponential
functions with constant growth rates (34, 71, 215). Gonz´alez-Est´evez et al. pooled
all worms together and therefore their data cannot discriminate between constant and
weakly size-dependent rates (71). Nevertheless, the value of 1.8 %/day for the degrowth
rate agrees well with our results in the considered size range of 1 − 5 mm2. Thomas
et al. tracked the growth of individual worms but did not explicitly analyze a size-
dependence of the rates (215). They obtain values below 5 %/day and most of them
lie in the range of 2 − 3 %/day, which is in good agreement with our data.
Oviedo et al. even claim to find linear functions for growth and degrowth with size
(150).
In consequence, the absolute values for growth and degrowth rates become
smaller in larger worms like in Fig. 5.5(d). Yet, otherwise the result does only roughly
agree with our data, mainly for the case of starvation, see Appendix E.2. Although
they even feed twice a week, they obtain very low growth rates, which are even lower
than their degrowth rates.
Finally, it has previously been claimed that nutritional status and not size determines
the growth dynamics (71). Yet, the respective measurements might not be able to
resolve this issue because it does not compare size matched worms of different feeding
histories.
In fact, our data with a controlled feeding history supports the opposite
hypothesis.
81
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.6.: (a) Hypothetical case of growth and degrowth dynamics with steady
state sizes (crosses). (b) Trends of the growth rates in terms of cell numbers for Smed
obtained from fits to the growth rates of Fig. 5.5(d) and the conversion curve to cell
numbers of Fig. 5.3(b). Note that gut stuffing might slightly change the obtained
values.
In summary, our measurements of growth dynamics agree in some cases quantitatively
and in others qualitatively with studies published by other groups but also go signifi-
cantly beyond. In particular, we were able to reveal the size dependence of the growth
rates by establishing a strict protocol of feeding and imaging and a highly accurate
image analysis. Furthermore, we were tracking individual worms and our measure-
ments span the full range in size in contrast to the previous works discussed.
Our analysis clearly shows that small worms show faster growth and degrowth dynamics
than large worms. Fig. 5.6 illustrates why this is a non-trivial observation. Naively one
could have assumed that there is one limiting quantity that makes the use of food less
efficient in larger worms, resulting in a decreasing growth rate with size, see Fig. 5.6(a).
Different amounts of food would result in a shift of the curve. In consequence, worms
would approach a steady state size, depending on the feeding frequency (crosses). In
contrast, the trends of the curves for Smed change between feeding and starvation
and the worms either grow or degrow. Fig. 5.6(b) shows our result in terms of cell
numbers. It suggests that there might be two different effects that depend in opposite
ways on the size of the worm and each of which dominates for either starvation or
maximum feeding. Stuffing effects as discussed in Sec. 5.2.2 might introduce correction
factors to the obtained growth rates. However, we expect no change in the qualitative
trends. In the next section, we explore three theoretical models that can account for
the size-dependence of the growth rates.
82
0246−505Cell number in 106Growth rate (cell number) in %/ day(a)(b) feeding every weekstarvationWorm sizeGrowth rateXXless feedingmore feeding5.3 Theoretical descriptions of cell turnover dynamics and energy flux
5.3. Theoretical descriptions of cell turnover dynamics and
energy flux
Fig. 5.6(b) shows nontrivial growth dynamics in flatworms depending on feeding con-
ditions. Growth corresponding to a change in cell number can be described by the two
processes of cell division and cell loss with effective rates kdiv and kloss, respectively:
N = kdivN − klossN .
With this, the growth rates plotted in Fig. 5.5(d) are given by
N /N = kdiv − kloss .
(5.1)
(5.2)
Thus, K = kdiv − kloss has to depend on N such that it decreases with N for feeding
and increases with N for starvation.
One possible scenario to explain this behavior is a change in the fraction of stem cells. In
fact, it has been reported that the neoblast fraction in Dugesia lugubris (110, 111) and
Dugesia tigrina (13, 16, 18) decreases as the worm grows. A higher stem cell fraction
in smaller worms would lead to a larger growth rate during feeding. Additionally, it
might be associated with a higher basal cell turnover, yielding a faster degrowth rate
in the absence of food. However, the difference in the neoblast fraction was estimated
to be less than 10%, which can not account by far for the observed size dependence
of the growth rates. Thus, it is rather unlikely that the fraction of neoblasts in Smed
varies by a factor of 2 or 3 like the growth rates. In fact, preliminary experimental data
obtained by the group of Jochen Rink suggests that the stem cell fraction stays rather
constant in animals of different sizes.
Feeding has a major effect on the growth dynamics. Therefore, we propose to consider
the metabolic energy Et in the worm as a limiting factor that influences cell proliferation
and cell death. At this point, we need to properly define this quantity, even though
the conclusion will be largely independent of the exact definition. For the rest of the
chapter, Et refers to the total amount of ATP and the ATP equivalent of any other
molecule storing metabolic energy like glucose and lipids, whether it is found inside
or outside the cells. Analogous descriptions can be found if only the freely available
energy outside the cells is considered or if one takes into account the abundance of
other limiting molecules like amino acids that need to be taken up by the food and that
are necessary for the worm to stay healthy and alive.
83
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
In a minimal model, the total energy Et increases due to net influx Jf by feeding and
decreases due to consumption by metabolic housekeeping with rate µ:
∂tEt = Jf − µN .
(5.3)
A typical value for the metabolic consumption per cell µ would be of the order of 10 pW
as estimated for Schmidtea polychroa (133). In humans with 4· 1013 cells and an energy
consumption of 107J/day (124), we estimate µ ≈ 1 pW. Further values obtained for
228 mammalian species from shrimp to elephant also show a decrease with size across
the range of 1 − 10 pW (240).
Eq. 5.3 assumes for simplicity that the energy stored in the dying cells can be completely
taken up by the remaining cells. We could consider an additional term of the form
−ρloss ec klossN , accounting for imperfect recycling upon cell death. Here, ec denotes
the metabolic energy per cell and ρloss is the fraction of it that is lost per dying cell.
The main conclusions of this chapter remain unchanged even for ρloss > 0.
We define the metabolic energy per cell as ec = Et/N and find for its dynamics
∂tec = jf − µ − K ec .
(5.4)
Here, jf = Jf /N is the net influx per cell and the last term describes a dilution effect.
As the worm grows, the same amount of energy has to be shared among a larger number
of cells.
In the following, we discuss three basic models on how the metabolic energy per cell
might effect cell division and loss. Given the simplicity of the models, they all fit
the noisy data reasonably well. Yet, they result in very different predictions about
the respective variables and parameters such as metabolic consumption and storage of
energy, which will be tested in future experiments.
The models describe three distinct scenarios of energy storage: (i) dynamic energy
storage, for which feeding increases the fraction of energy stored in the worm, (ii)
energy storage of fixed proportion, for which the worm is not able to store additional
amount of energy upon feeding, and (iii) size-dependent energy storage, for which the
fraction of energy stored depends on the size of the worm and not explicitly on feeding.
In the first model, the metabolic energy per cell ec changes as a direct read-out of size
and feeding conditions and directly regulates cell division and cell loss. In the second
and third model, the metabolic energy is quickly regulated to a physiologically preferred
target value e0 by adjusting the cell turnover rates. Thereby, the mechanism implicitly
draws on the idea of integral feedback control, which is commonly used in engineering
applications to drive a dynamic variable to a pre-set value.
84
5.3 Theoretical descriptions of cell turnover dynamics and energy flux
5.3.1. Model 1: Dynamic energy storage
In a most simple model for metabolic growth control, the cells are assumed to directly
respond to the amount of metabolic energy per cell ec. If there is an excess in energy,
cells divide more often, and if there is a lack of energy, the death rate increases. Let us
consider a simple case, for which the growth rate depends linearly on ec:
K = K0 (ec/es − 1)
(5.5)
with the constant parameters K0 and es, which determine the rate of growth and the
energy, for which the worm switches between growth and degrowth, respectively.
The change in energy ec as a function of ec is plotted in Fig. 5.7(a). During starvation
periods, when jf = 0, the growth rate is decreasing, which requires ec < 0 (red curve).
The maximum of ec is at ec = es/2 and from ec < 0 follows that µ > esK0/4.
During feeding with jf > µ, the curve in Fig. 5.7(a) is shifted upwards and ec ends up
in a regime, for which the organism grows (blue). If jf was constant, there would be
a stable steady state with e∗
c > es
for jf > µ and the system is in the growing regime. In order for the growth rate to
decrease with worm size as seen in Fig. 5.6(b), the influx due to feeding jf (N ) must
not be constant but has to be a decreasing function of N .
c = es/2 +(cid:112)(es/2)2 + (jf − µ)es/K0. We see that e∗
Fig. 5.7(b) illustrates a time course of N , when going through several rounds of feeding
and starvation, always switching at a certain size (horizontal lines). Especially in the
beginning of the starvation interval, we see an overshoot, for which the worm still
grows although the feeding has stopped. Nevertheless, we want to stress that the
growth and degrowth behavior is rather generic. The growth and degrowth rates in
Fig. 5.7(c) collapse and show the same size-dependence, irrespectively of the initial
values for energy and cell number. Any perturbation decays quickly and there is no
strong dependence on feeding history.
We can fit this model to the experimental data, see Fig. 5.8(a). The plots shows the
size dependence of the influx jf and illustrate that the metabolic energy per cell ec acts
as a read-out for both feeding conditions and size at the same time.
85
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.7.: Model 1 assumes that cell division and cell death directly depend on
the metabolic energy per cell ec, which represents a dynamic energy store. (a) During
starvation, ec and thus the degrowth rate decreases. During feeding, ec approaches
the growing regime. Note that e∗
√
c > es is not a steady state of the total system
N ) decreases with N . (b) Time course of the cell
because the influx jf (N ) = j0/(1 +
number when switching between feeding (blue) and starvation (red) for N = 0.5 · 106
and N = 4.5 · 106, respectively (dashed lines). (c) We observe a generic dynamics,
irrespective of the initial cell number, energy or feeding scheme.
86
0growthdegrowthfeedingstarvation02004006008001000012345Time in daysCell number in 106(a)(b)(c)Varying the intial cell number:Varying the intial energy per cell:Varying the switching between feeding and starvation:−202412345Cell number in 106Growth rate in %/day0−202412345Cell number in 106Growth rate in %/day0−202412345Cell number in 106Growth rate in %/day05.3 Theoretical descriptions of cell turnover dynamics and energy flux
5.3.2. Model 2: Fixed proportion energy storage
We now assume that the energy per cell ec is quickly regulated to a physiologically
preferred target value e0. In this scenario, there might still be specialized energy stores
like fat cells but they are only formed proportionally to the worm size and not in
response to feeding. Any metabolic energy outside the cells is also quickly regulated,
which is typical for the metabolism of animals. For example, sugar levels in the human
blood are under tight regulation by the insulin pathway and insulin-like molecules
are found even in the simplest unicellular eukaryotes (114). Yet again in contrast to
humans, here we assumed that no additional long-term storage cells are formed upon
feeding. Instead, the metabolic energy is regulated by adjustment of cell division and
cell death. We can picture an integral feedback scheme, according to which K changes
depending on the difference between ec and e0:
τK K = K0(ec/e0 − 1)
(5.6)
This control theoretic approach substantially differs from the case of model 1 in Eq. 5.5.
Given a fast regulation to the target value e0 (τK → 0), K relaxes to
K∗ =
jf − µ
e0
,
(5.7)
which defines the nullcline of ec = 0 according to Eq. 5.4.
Fig. 5.6(b) shows that the degrowth rate is decreasing during starvation, where jf = 0.
Thus, µ has to increase for smaller worms. Fig. 5.8(b) illustrates the size-dependence of
µ based on our measurements. This can be interpreted in the sense that the metabolism
is less efficient in smaller animals, which is in agreement with the typical metabolic sca-
ling laws found across the animal kingdom (7, 190) and which have also been observed
in flatworms (71, 92, 133).
During feeding the growth rate decreases with size. Thus, jf has to decrease with N
and even has to overcompensate for the opposite effect of µ. This suggests that food
uptake is less efficient in larger animals, see Fig. 5.8(b).
Taken together, we assumed that the metabolic energy is the limiting factor determined
by the corresponding nutrition influx and consumption, to which turnover rates are
adjusted accordingly. We obtained two functions, one for µ and one for jf , which
depend on worm size in opposite ways and therefore can account for the opposite
trends in the growth and degrowth dynamics.
As a side note, one might question what is cause and what is consequence. While
it is likely that the metabolic energy is the limiting factor during starvation periods,
87
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.8.: (a) Model 1 is characterized by energy stores that are filled and depleted
depending on feeding and worm size. The stored energy per cell ec as well as the
metabolic rate µ and the feeding influx jf are obtained by minimizing the variance
between the experimental data and numerical solutions of our model. The values
are given relative to the energy es, for which the worm switches between growth and
degrowth. (b) Model 2 draws on energy stores with fixed proportions. We compute µ
and jf relative to the energy per cell from the fits to the growth and degrowth rates
in Fig. 5.5(d). The energy per cell ec = e0 is assumed to be constant. (c) In model 3
with size-dependent energy stores, the energy per cell ec changes in a size-dependent
manner, while µ = µ0 and jf are constant. Degrowth and growth rates are fitted by
the same function with a different amplitude corresponding to the feeding influx. (d)
Basic scheme of model 3: During feeding, small worms invest their metabolic energy
in growth, yet large worms store an increasing fraction of the available energy. During
starvation, large worms degrow more slowly as they can deplete their energy stores.
88
02468012345Cell number in 106Feeding influx ( )01234500.511.52Cell number in 106Metabolic rate ( )02468012345Cell number in 106Feeding influx ( in %/day)02468101214012345Cell number in 106Feeding influx ( in %/day)01234500.511.52Cell number in 106Energy per cell ( )0123450123Cell number in 106Energy per cell ()01234501234Cell number in 106Metabolic rate ( in %/day)(b) Model 2: Fixed proportion energy storage012345−4−20246Cell number in 106Growth rate in %/ day(c) Model 3: Size-dependent energy storage012345−4−20246Cell number in 106Growth rate in percent / dayfeeding of small worms:storagegrowth++feeding of large worms:storagegrowthstarvation of large worms:storagemaintenance+012345050100150Cell number in 106Energy per cell ( in days)(d) Energy storage in model 301234500.511.52Cell number in 106Metabolic rate ( in %/day)(a) Model 1: Dynamic energy storage012345−4−20246Cell number in 106Growth rate in %/dayfeedingstarvation5.3 Theoretical descriptions of cell turnover dynamics and energy flux
this is not necessarily the case for maximum feeding. In the presence of an abundance
of food in the gut, cell division response could rather be limited by other factors as
the maximum rate of DNA replication and the available number of stem cells. This
would invert the argument: there would be a generic, yet size-dependent response of
cell division (and maybe cell death) and the size dependence of the net influx jf is
merely a consequence of it.
5.3.3. Model 3: Size-dependent energy storage
In the scenario above, we introduced a constant target value e0 for the metabolic energy
that determined the turnover rates. In our third and final model, we again consider
a quick relaxation to such a target energy, but now this energy e0(N ) changes with
worm size. For example, the fraction of energy-rich cells could increase in large worms.
We again obtain Eq. 5.7 but now µ and jf are assumed to be constant for simplicity,
while e0 depends on N . The three variables are plotted in Fig. 5.8(b). Even with only
one size-dependent quantity, we obtain a reasonable fit to the data.
If additionally
µ was allowed to change as suggested by the measurements of metabolic scaling laws
(7, 92, 133, 190), the fits would even improve.
The basic logic of such a mechanism based on energy storage cells is shown in Fig. 5.8(d).
While small worms invest all available energy into growth, large worms store part of
their energy uptake and thus grow more slowly. In consequence, large worms deplete
these energy stores during starvation and initially degrow more slowly. Importantly,
we assume that the size of the stores scales with worm size. Thus, only an effective
size-dependence of growth and degrowth is observed, but no explicit dependence on
feeding history.
5.3.4. Discussion of the turnover models
We have explored three models that are based on very different assumptions and dif-
fer greatly in the microscopic details, yet they all can account for the observed size-
dependent growth behavior. This is an interesting observation, which one should gen-
erally take into consideration when comparing microscopic models to more macroscopic
data. A theoretical model is typically associated with a specific coarse-graining level.
As a necessary condition, it needs to agree with observations on larger scales, but this
is not sufficient for the model to be unique. In order to distinguish between several
microscopic models, one usually needs measurements on the same level.
89
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
In our case, the theory makes specific testable predictions about the influx, consump-
tion and storage of energy, as illustrated in Fig 5.8. In close collaboration with our
experimental colleagues, we have planned further experiments to measure these quan-
tities and to parametrize the models. First, it needs to be verified that the fraction of
stem cells is approximately constant in Smed and cannot explain the growth dynamics.
Furthermore, our collaborators are conducting experiments at the time of writing to
determine metabolic rates and feeding influx. Additionally, they aim to identify po-
tential stores for the metabolic energy and probe the effects of inhibiting them. The
existence of fat cells has already been reported for other flatworm species (93).
Fig. 5.4 already shows that the stuffing peak is approximately proportional to worm size.
If therefore the average amount of food that becomes available per cell is independent
of the size of the worm, it will support the third model. Furthermore, we have discussed
in Section 5.2.1 that the scaling laws for worm mass and total protein mass might also
be in agreement with an increased storage of lipids or glycogen in larger worms.
It
remains to be answered whether the scaling of metabolic rates and storage cells can
explain the growth and degrowth behavior in Smed .
Even though the models are highly simplified and gut stuffing might introduce cor-
rections, we can extract order of magnitude estimates from the fits to the data. For
example, we can estimate from all three models that each cell stores on average as much
energy as needed to maintain it for approximately ec/µ = 50 to 150 days. In compari-
son, the metabolic rate in humans is µ ≈ 1 pW and the total energy stored in human fat
cells (as the main energy storage) is 4 · 108 J (124). Thus, ec = 10−5 J, which amounts
to a similar ratio of ec/µ = 130 days. Furthermore, based on measurements in the flat-
worm Schmidtea polychroa and on the typical metabolic scaling laws (124, 133, 240),
we expect the metabolic rate of Smed to be in the range of µ ≈ 10 pW per cell. Thus,
according to our models, each cell stores as much as ec ≈ 10−5 to 10−4 J. Note that
this estimate is based on the assumption of perfect recycling of the energy upon cell
death. For imperfect recycling, the value is reduced.
Metabolic energy can be stored in various ways: in lipids or glycogen, in specialized fat
cells or distributed among many different cell types. An interesting observation in this
respect is the presence of germ line precursor cells in asexual flatworms. Cells without
a purpose should vanish during evolution. Thus, we might speculate that the germ
line precursors in asexual worms might have an additional function for the metabolic
housekeeping. In analogy to the third model, they are especially prominent in large
worms and can potentially serve as energy stores during starvation.
90
5.4 Control logic for cell turnover and growth
5.4. Control logic for cell turnover and growth
5.4.1. Measuring cell turnover on various scales
So far, we have investigated how the cellular behavior might depend on feeding condi-
tions and worm size but we have not explicitly distinguished between the adjustment of
cell division on the one hand and cell loss or death on the other hand. In this section,
we now discuss how these individual processes might be controlled. In order to unravel
the control logic, we propose experiments on various scales, ranging from the level of
individual cells to a whole tissue and to the averaged turnover dynamics across the
entire worm. The main questions are:
• Does feeding and worm size affect rather cell proliferation or cell death?
• How do both processes influence each other?
• To what extent does aging play a role on the cellular scale?
The first question will be addressed mainly at the cellular level and the second and
third question at the tissue level.
Turnover dynamics at the cellular level. -- Previous measurements indicate that there
is an catabolic default state during starvation periods, in which cell loss dominates over
cell division. Upon feeding, there are short-term mitotic bursts leading to growth, as
sketched in Fig. 5.9 (12, 13, 16, 71, 101, 139, 142, 170). However, it is not clear whether
the cell loss rate also changes as part of the feeding response and to what extent division
and loss depends on feeding history. It has been suggested that apoptosis rates might in
fact increase after the flatworms have been fed, especially after long starvation periods,
as a means of cell renewal (71, 159).
Assessing these questions is complicated by the fact that cell loss is poorly defined.
Apart from the classical apoptosis pathway many other processes might contribute to
cell loss such as other mechanisms of programmed cell death but also uncontrolled
shedding of epidermis cells.
In contrast, cell division can be well characterized.
In
Appendix E.3 we discuss a measurement scheme to extract division rates. By comparing
cell division to worm growth, we will also be able to determine to what extent cell death
rates are regulated.
Turnover dynamics at the tissue level. -- By monitoring the turnover dynamics in
epidermis cells, we can exemplify how cell addition and removal work together to result
91
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.9.: A hypothetical model that comprises (i) an catabolic default state, for
which the cell loss rate (dashed green) is larger than the cell division rate (solid red),
and (ii) short proliferation responses upon feeding.
in a well controlled size regulation.
In the next section, we introduce a respective
measurement protocol. We show preliminary data, which suggests that cell death
depends on the age of the cells.
We also discuss further experiments to confirm the effect of aging and to investigate
how the individual turnover processes influence each other. On the one hand, turnover
could be a rather stochastic process, in which cells are inserted in the tissue and deleted
independently, yet on average at balanced rates. On the other hand, insertion and
deletion could be tightly linked such that a cell is preferentially removed in the presence
of a new cell that replaces it. We suggest a double labeling experiment which can
potentially distinguish the two control paradigms.
Turnover dynamics at the organismal level. -- After constructing a comprehensive
model on how turnover and growth is regulated based on the experiments on the cellular
and the tissue scale, we can validate the model by measurements across the entire
worm. To this end, we adapt a protocol that has originally been developed to analyze
cell turnover in human brain cells (203). In Appendix E.4, we provide the theoretical
framework to set up the experiment.
5.4.2. Analyzing turnover of the epidermis as an example tissue
Epidermis cells as the outermost skin cells can be non-invasively labeled by soaking
the worms in CFSE solution (carboxyfluorescein diacetate succinimidyl ester). It has
been observed that the labeled cells disappear within days as a result of turnover of
the epithelial tissue. From this experiments, we can estimate turnover rates of the
92
Loss / Division rateTime(i)(i)(ii)5.4 Control logic for cell turnover and growth
tissue and investigate whether older cells are more prone to die. We will also propose
a protocol with two consecutive labeling pulses, which might enable us to understand
how insertion and deletion of cells is coordinated.
5.4.2.1. Measuring turnover rates by single pulse labeling
Cell turnover processes. -- We describe turnover in a cell population by three different
processes: (i) deletion of cells, (ii) insertion of cells and (iii) replacement of cells (i.e.
coupled insertion and deletion), see Fig. 5.10(a). Note that deletion potentially includes
many different processes: the induced removal by long range signals or quorum sensing
in the tissue, cell-autonomous decisions as well as uncontrolled shedding.
In order to measure turnover dynamics, we labeled all cells in the epidermis by soaking
in CFSE and counted the fraction of labeled cells in the field of view after time t, see
Fig. 5.10(b)-(c). The experiments were performed by Sarah Mansour in the group of
Jochen Rink. The author analyzed the data by automatically counting the cells with a
custom MATLAB code.
Dynamics of a labeled cell population. -- Let us consider the number density n(a) of
cells of a certain age a. The total number of cells N in the tissue and the number of
labeled cells N(cid:96) are given by
(cid:90) ∞
(cid:90) ∞
N =
n(a) da ,
N(cid:96) =
n(a) da .
(5.8)
Due to cell turnover, these quantities change in time as follows
0
t
N = (ki − ¯kd) N
N(cid:96) = −(¯kd(cid:96) + ¯kr(cid:96)) N(cid:96) .
(5.9)
(5.10)
(cid:90) ∞
Here, we consider the rate for cell insertion ki and average rates for the replacement of
labeled cells ¯kr(cid:96) as well as for the deletion of labeled and unlabeled cells ¯kd(cid:96) and ¯kd:
¯kr(cid:96) =
kr(a) n(a)
t
N(cid:96)
da ,
¯kd(cid:96) =
kd(a) n(a)
t
N(cid:96)
da ,
¯kd =
kd(a) n(a)
0
N
(cid:90) ∞
(cid:90) ∞
da .
(5.11)
(5.12)
When taking snapshots of only part of the tissue, we do not obtain absolute numbers
but the fraction of labeled cells in the field of view Φ(cid:96) = N(cid:96)/N , which obeys
Φ(cid:96) = −(cid:0)¯kr(cid:96) + ki + (¯kd(cid:96) − ¯kd)(cid:1) Φ(cid:96) .
The fraction of labeled cells might not only decrease due to replacement (first term),
but also due to a dilution effect as unlabeled cells get inserted into the tissue (sec-
ond term). Furthermore, the third term arises if labeled cells and unlabeled cells are
93
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
Figure 5.10.: (a) We distinguish three processes of cell turnover. (b) Labeled epider-
mis cells (CFSE staining - green) disappear within a few days due to turnover. Nuclei
are labelled in blue (Hoechst staining). (c) Measurement scheme of the CFSE labeling
experiment. (d) Fraction of labeled cells at a certain time after the label is applied. We
fit an exponential curve with a constant tissue turnover rate ktis = 0.27/day (dashed
blue) as well as linear function corresponding to a fixed life time of the cells amx = 7.8
days (red). (e) Measurement scheme with irradiated (stem cell depleted) worms. (f)
For T = 1 day, the turnover in irradiated worms (green) compares to non-irradiated
worms (black). Yet, for T = 5 days after irradiation, the labeled epidermis cells do
not vanish from the tissue anymore (red). All measurements were performed by Sarah
Mansour in the group of Jochen Rink. The author analyzed the data and fitted the
models.
94
(a)(b)(i) deletion(ii) insertion(iii) replacementas independent processesas tightly coupleddeletion & insertion(c)labelingmeasuringTime t(e)(d)(f)irradiatingmeasuringTtlabelingTime t in days0246Label fraction00.20.40.60.8101234Time t in daysLabel fraction00.20.40.60.815.4 Control logic for cell turnover and growth
deleted at different rates (e.g. in an age-dependent manner). In contrast, if cells are
deleted stochastically (e.g. according to a Poisson process), kd(a) is constant and the
corresponding term vanishes: ¯kd(cid:96) − ¯kd = 0.
Feeding conditions and the sizes of the worms are approximately constant during the
course of the experiment. Thus, if the turnover does not depend on the age of the cells
(i.e. kd and kr are constant), we can expect an exponential solution Φ(cid:96) = Φ(cid:96),0 e−t ktis to
Eq. 5.12 with a constant rate ktis = kr + ki. Yet, the preliminary data in Fig. 5.10(d)
does not strongly support an exponential law (dashed blue).
Another limiting case is that cells have a fixed live span and only die at age amx. Thus,
the rates of deletion and replacement are kd(a) = δ(a − amx) and kr(a) = δ(a − amx).
Note that the worms only grow and degrow by a few percent per day, such that we can
assume that the tissue size only changes very little in comparison to the turnover time
scales of a few days. In consequence, the age distribution n(a) will be approximately
homogeneous and the solution to Eq. 5.12 a linear function Φ(cid:96) = Φ(cid:96),0 − t/amx. This
simple linear model (solid red) appears to agree better with the data in Fig. 5.10(d).
In fact, the truth might be in between the two limiting cases. This would resonate with
an age-dependent turnover of the cells in the tissue such that old cells are more prone
to deletion or replacement.
The time dependence enters because the age distribution of labeled cells is shifted over
time. In order to extract in what way the cellular age influences deletion or replacement
and which of the two processes is mostly affected, we need more sophisticated experi-
ments as discussed below. Also note that the presented data is only very preliminary
and has to be confirmed by further measurements.
Labeled cell fraction in irradiated worms. -- We briefly comment on a second pilot
experiment by Sarah Mansour, see Fig. 5.10 (e)-(f). Here, the measurement started
at varying times T after the worms were depleted of all stem cells by γ-irradiation.
First, we observe that tissue turnover is not directly affected by the loss of stem cells:
for small T the label fraction (green) decreases in the same way as for non-irradiated
worms (black). Labeled cells are replaced by progenitor cells, which were not deleted by
irradiation. Thus, there seems not to be a long-range signaling effect from the stem cell
pool. In contrast, T = 5 days after irradiation (red), the fraction of labeled cells does
not decrease anymore. In conclusion, dividing cells (which are affected by irradiation)
or corresponding signals need 5 days to reach the epidermal tissue.
95
5. QUANTIFYING FLATWORM GROWTH AND CELL TURNOVER
5.4.2.2. Monitoring cell turnover by double pulse labeling
We propose a double-labeling experiment using a second labeling pulse of a different
marker (DDAO-SE) applied after a short time interval ∆t, see Fig. 5.11(a). As a result,
there are two labeled cell populations -- cells with both labels N2 and young cells with
(cid:90) ∞
only the second label N1:
(cid:90) t+∆t
N1 =
n(a) da ,
N2 =
n(a) da .
(5.13)
t
t+∆t
Similar to Eq. 5.12, the ratio Φ12 = N1/N2 obeys
Φ12 = −(cid:0)(¯kd1 − ¯kd2) + (¯kr1 − ¯kr2)(cid:1) Φ12 ,
(5.14)
which includes the deletion and replacement rates of both labeled cell populations.
Additionally, we propose an irradiation experiment as depicted in Fig. 5.11(b). For this,
we choose the time interval T such that there are still new cells arriving in the epidermis
during ∆t but not so after the second labeling pulse. According to our measurements
with a single label in Fig. 5.10(f), T + ∆t ≈ 5. In consequence, (¯kr1 − ¯kr2) = 0 during
the time interval t:
Φ12,irr = −(¯kd1 − ¯kd2) Φ12 .
(5.15)
This shows that Φ12,irr only changes for age-dependent deletion rates. Thus, by com-
paring the dynamics of Φ12 in non-irradiated and Φ12,irr in irradiated worms given by
Eq. 5.14 and 5.15, respectively, we can deduce whether deletion or replacement depends
on the age of the cell, as illustrated in Fig. 5.11(c). If Φ12,irr changes according to a
time-dependent rate, deletion is related to the age of the cells.
If instead only the
dynamics of Φ12 are determined by a time-dependent rate, replacement relies on the
age of the cells.
Furthermore, we can extract actual rates for the individual processes by comparing the
results to the previous experiments. For example, ¯kd2 + ¯kr2 in Eq. 5.14 is equivalent to
¯kr(cid:96) + ¯kd(cid:96) of the experiment with a single label. By combining these measurements, we
can find the functional relationship for how deletion and replacement changes with the
age of the cells.
The same experiments can be performed in starving and well fed animals to reveal
the connection between feeding status and turnover rates. One hypothesis would be
that there is no insertion during starvation, yet replacement dominates and the tissue
slowly shrinks due to a small deletion rate. After feeding, insertion leads to growth of
the tissue.
96
5.5 Summary
Figure 5.11.: Double pulse labeling scheme in (a) non-irradiated and (b) irradiated
animals. (c) The double labeling experiment allows us to deduce whether deletion or
replacement depend on the age of the cells.
5.5. Summary
Our aim is to study cell turnover and growth in a comprehensive way, bridging the
scales from a single cell to the average behavior at the level of the organism. For this,
we have carefully designed several experiments to determine the flatworm size and to
characterize its growth dynamics. In particular, we developed a protocol for precise
measurement of the worm area based on the analysis of movie sequences with a custom
MATLAB routine. The data was compared to previous studies, which in parts yielded
controversial results.
We observed that the growth and degrowth rates are dependent on worm size. Smaller
worms grow and shrink faster. This resonates with the picture that small worms might
be more juvenile with a potentially higher turnover rate and a different metabolism
(71, 111). Based on the idea that the available metabolic energy is the main deter-
minant for growth, we have developed three distinct classes of microscopic models,
which can all be fitted to the macroscopic growth data. The models describe different
paradigms of energy storage and make specific predictions that are tested by our col-
laborators in current experiments.
In particular, they also differ in the way energy
inflow (i.e. digestion of food) and energy outflow (i.e. metabolic consumption) varies
with size, which directly relates to the question about the limits to growth.
Finally, we propose further experiments, which can unravel the control logic of cell and
tissue turnover. We provide the theoretical framework and discuss first preliminary
data suggesting an age-dependent death of cells. The experiments will be performed in
the laboratory of Jochen Rink by the author and other members of the group.
97
(a)irradiatingmeasuringTt1. labeling2. labelingΔtmeasuringt1. labeling(CFSE)2. labeling(DDAO-SE)Δtwild-typetime interval tno age-dependenceonly replacement is age-dependentonly deletion is age-dependenttime interval ttime interval tage-dependecies of replacement & deletion add uptime interval ttime interval tirradiatedage-dependecies of replacement & deletion cancel(b)(c)6. Summary and outlook
In this thesis, we analyzed various aspects of growth and body plan scaling, drawing
particular inspiration from the flatworm Schmidtea mediterranea (Smed ) as a model
animal. For this purpose, we combined theoretical descriptions and analysis of experi-
mental data.
As a first approach, we analyzed previously proposed mechanisms for self-organized
pattern formation as well as pattern scaling in Chapter 2. We systematically extracted
the requirements for scaling and demonstrated that they are challenged by the fas-
cinating regeneration and re-patterning capabilities of flatworms. On this basis, we
presented a minimal model for fully self-organized and self-scaling pattern formation
in Chapter 3. This mechanism is capable of spontaneously generating concentration
gradients and expression regions of involved molecules that robustly adjust to the size
of the system.
In the future, it will be interesting to further compare the theoretical framework and
the derived predictions to the Wnt/β-catenin system, which is associated with head-tail
(AP) polarity in Smed (4, 9, 81, 82). Preliminary measurements by our collaborators
indicate that the β-catenin gradient as well as expression profiles of various Wnts in
fact scale with worm size. Our theory enables us to interpret the data and to suggest
additional experiments to unravel the mechanism of scaling in flatworms. Further
beyond, the results might also enrich the discussions on signaling gradients in other
systems, such as the developing fly wing, for which it is still a matter of debate how
scaling is established (11, 26, 55, 233).
Smed only recently evolved to become a model organism, thus the basic mechanisms for
axis specification are far from being well understood (117). It is known that components
of the Wnt/β-catenin system are especially present in the tail (4, 9, 81, 82). However,
there could be also an additional patterning system originating from the head or even
a compartmentalization by several systems along the body axis. If there are several
anterior-posterior (AP) polarity systems, how do they interact? And to what extent is
the patterning self-organized or relies on pre-existing cues such as a polarized tissue or
wound-specific signals after amputation? We are currently analyzing a large data set
98
comprising the spatial changes in gene expression after RNAi treatment and expression
time courses during regeneration in order to answer these questions.
Perpendicular to the AP axis, there are also two more body axes in dorsal-ventral and
medial-lateral direction (117). How do the patterning systems for the different body
axis influence each other? Is there a hierarchy of axis formation or do the axes emerge
simultaneously? One particular curious case is the formation of a straight midline
as an important signaling center. The experimental data suggests that a repulsive
signal from the body margin positions the midline (3, 117, 121). Yet, is it sufficient to
account for this narrow, straight row of distinct cells, which tends to re-emerge from
the anterior and posterior poles during regeneration? As a complementary approach,
we have started to design a model explaining midline formation on the basis of cell
proliferation and resulting cellular flows in the tissue.
Besides these long-range patterning mechanisms that are classically assumed to rely on
secreted, motile molecules, there are also other polarity cues. For example, the planar-
cell-polarity (PCP) system is based on direct cell-cell interactions and results in cellular
polarisation consistently across the tissue. In order to determine the coupling between
both mechanisms, we have started analyzing the motility of flatworms, when various
Wnt-pathway components have been knocked out by RNAi feeding, see Chapter 4.
Resulting movement phenotypes report on a dysfunctional cilia carpet associated with
an impaired PCP system.
For a further, more detailed investigation of worm motility, we adapted principal com-
ponent analysis to apply it to the highly deformable worm body. We could demonstrate
that during normal gliding motion worms steer their path by bending in the preferred
direction of movement. Additionally, we provided the first quantitative account of an
alternative motility mode, called inchworming. This appears to be a tightly controlled
behavioral response to impaired cilia functionality with a characteristic frequency of
about 1/4 Hz. Such stereotypic behaviors have been previously described in the much
simpler nematode C. elegans (207, 208, 209). It is a fascinating observation that more
complex organisms such as flatworms still show very generic motility patterns, which
might have emerged during evolution as an optimized strategy and are expected to
relate to the structure of the muscular plexus and the nervous system. Similarly, we
were able to demonstrate the applicability of shape mode analysis to the variable head
shapes between different species. We are planning to extend this analysis to many more
species of the large flatworm collection in the laboratory of Jochen Rink. As both, spe-
cific morphologies and specific movement strategies, are expected to emerge due to the
99
6. SUMMARY AND OUTLOOK
evolutionary pressure in the respective environmental niche, our analysis provides the
basis for future research relating form and function.
In the last chapter, we quantified growth and turnover dynamics in flatworms. Interes-
tingly, we obtained size-dependent growth rates. Small worms appear to grow and
degrow faster than large worms. This particular growth and degrowth behavior of the
worms is the coarse-grained result from the underlying processes of cell division and
cell loss. Thus, we devised several microscopic models describing specific rules on the
cellular scale to explain the observed growth dynamics on the macroscopic level of the
organism. The models are based on the idea that there is a limiting quantity that
"Nothing in biology makes sense except in the
light of evolution." -- ditto, Theodosius
Dobzhansky, 1973 (41)
(i) is provided by feeding, (ii) is permanently con-
sumed by the cells and (iii) affects the division
and loss rates of cells. As an example, we expli-
citly consider the metabolic energy as the limiting
quantity. All of our three models fit the growth data equally well. Importantly, each
model is based on distinct assumptions on how the energy is stored and how the energy
availability affects the division and loss rates. Furthermore, each model makes specific
predictions about how the three considered processes of energy influx by feeding, energy
storage and energy consumption have to depend on the size of the worm to account for
the measured growth data.
With this result in mind, we contribute to the design of future experiments that will
enable us to distinguish between the models. Thereby, we aim to bridge scales by
providing an explanation for the organismal growth dynamics on the macroscopic level
in terms of cellular behavior on the microscopic scale. The outcome might also hint at
systemic limitations to growth. For example, if the food uptake is size-dependent and
relatively decreases with worm size, there will be an upper limit at which the worm
is not capable to sustain growth anymore.
It then poses the question whether this
is a physical limit set by constraints of the worm body or whether evolution has not
selected for a more efficient uptake because worms anyways do not grow bigger for
other reasons. Similarly, energy stores that depend on either size or feeding will trigger
further investigations on the specific cells that store the energy. Are there specialized
fat cells or does a broad range of cells store the energy? An interesting idea is related
to the formation of the germ line. For sexual flatworms, there might be a tradeoff
between growth and the development of a reproductive system. Small worms would
rather grow to survive, large worms would rather invest in reproduction. Yet, germ
line precursor cells still exist in the asexual strain used in the experiments. From an
100
evolutionary viewpoint, this might hint at a dual role of these cells, suggesting that
they are involved in growth control or energy storage.
The level of detail in the theory reflected the coarse-graining level of the experimental
data. In a next step, we aim to go beyond and include further measurements to dissect
the control logic of growth and cell turnover. Which of the two processes of cell division
and loss is affected by feeding status and worm size in which way? How do the two
processes influence each other? To what extent are these turnover processes related to
aging of the cells as well as aging of the organism? There are many open questions and
we started looking into some of them. For example, we have discussed a theoretical
framework for measurements from the cellular to organismal scale and also analyzed
preliminary data, which suggests an age-dependent replacement of cells. Eventually,
we aim to develop a comprehensive picture of how cell turnover and organism growth
is controlled.
This thesis addresses various aspects of growth, cell turnover and scalable body plan
patterning in flatworms. It illustrates the sheer diversity of biological processes, which
are fascinating for biologists and physicists alike. Within a collaboration between ex-
perimentalists and theoreticians, we applied physical concepts to quantitatively analyze
and interpret experimental data. Importantly, the gained insights enabled us to ask
new questions and to suggest future experiments, which will ultimately help to grow
our general understanding of homeostasis and growth, development and regeneration
in multicellular organisms.
101
A. Reaction-diffusion systems:
fixed points and scaling
A.1. Morphogen dynamics with linear degradation
A.1.1. Reaction, diffusion, advection and dilution
Here, we consider general morphogen dynamics in a growing tissue. We assume that
the morphogens are produced by a source term ν = ν((cid:126)r) and spread in the system with
an effective diffusion coefficient D while being subject to linear degradation with rate
β. The dynamics can be described by the following convection-diffusion equation for
the morphogen concentration C = C(t, (cid:126)r) (11, 31)
∂tC = D ∇2 C − β C + ν − ∇(cid:0)(cid:126)u C(cid:1) .
(A.1)
The last term corresponds to dilution and convection due to tissue growth with a
velocity field (cid:126)u = (cid:126)u(t, (cid:126)r).
Spreading of morphogens from a localized source results in graded concentration pro-
files. Many systems can be considered to be homogeneous with respect to all but one
direction (e.g.
the x-direction).
In this case, the graded profiles C = C(t, x) only
depend on the respective spatial coordinate and the differential equation simplifies to
∂tC = D ∂2
= D ∂2
x C − β C + ν − ux∂xC − C∇(cid:126)u
x C − (β + ∂yuy + ∂zuz) C + ν − ∂x(uxC) .
(A.2)
Note that dilution by growth in the directions perpendicular to the graded profile can
effectively be described by a degradation term. For slow growth dynamics, we can
neglect the convection-dilution terms and recover Eq. 2.1.
A.1.2. Steady state solution neglecting tissue growth
The steady state solution of Eq. A.2 can be found in (31). Yet, often the effect of tissue
growth is considered to be negligible or a second order perturbative effect to the solution
of Eq. 2.1. Thus, let us consider this limit of slow tissue growth. Furthermore, D and
102
A.1 Morphogen dynamics with linear degradation
β are assumed to be constant in space across the tissue and constant in time within
the time scale of the morphogen dynamics. As a result, the morphogen concentration
approaches a steady state profile of the form
C∗(x) = c1 exp(x/λ) + c2 exp(−x/λ) .
The steady state solution for reflecting boundary conditions is given by
for 0 ≤ x < w
for w ≤ x ≤ L
sinh(L/λ) cosh(L/λ − x/λ)
C∗(x) =
cosh(x/λ)
α
β
In the limit of a small source and a large system (w (cid:28) λ, λ (cid:28) L), the steady state
simplifies to
sinh(L/λ)
sinh(w/λ)
1 − sinh(L/λ−w/λ)
(cid:40)e−x/λ
C∗(x) ≈ α w√
D β
2 e−L/λ
for x (cid:28) L
for x = L .
(A.3)
(A.4)
(A.5)
(A.6)
(A.7)
(A.8)
The square-root of D and β in the amplitude also emerges if there is no degradation in
the source region, see Appendix A.1.4.
A.1.3. Relaxation to the steady state
The degradation rate defines the time scale of relaxation to the steady state of Eq. 2.1.
Let us consider a concentration C(t, x) = C∗(x) + c(t, x) that deviates from the steady
state by a perturbation c(t, x). The dynamics of c(t, x) are described by
∂t c = D ∂2
x c − β c .
The concentration in Fourier space c = c(t, s) obeys
∂t c = −(cid:0)(2πs)2D + β(cid:1) c ⇒ c = c,0 e
−(cid:0)(2πs)2D+β(cid:1)t
and the back-transform yields (see Eq. B.7 for details)
c(t, x) = e−βt
dx(cid:48) c(0, x(cid:48))
−(x(cid:48)−x)2
4Dt√
e
4πDt
.
(cid:90) ∞
−∞
Besides the term for diffusive spreading that homogenizes the initial perturbation
c(0, x), there is an exponential damping term. Thus, the perturbation of the steady
state C∗ decays on a time scale of 1/β.
103
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
A.1.4. Steady state without morphogen degradation in the source
The reaction-diffusion equation without morphogen degradation in the source region is
∂tC = D ∂2
x C − β C Θ(x − w) + α Θ(w − x) .
(A.9)
With reflecting boundary conditions as in Eq. 2.2, we obtain
C∗(x) =
√
√
α w
Dβ tanh(L/λ−w/λ)
α w
Dβ sinh(L/λ−w/λ)
2D (w2 − x2)
+ α
cosh(L/λ − x/λ)
for 0 ≤ x < w
for w ≤ x ≤ L
(A.10)
While we obtain a power law inside the source, the expression outside the source is
identical to the limit w (cid:28) λ of Eq. A.4.
A.2. Gradient scaling with expander
A.2.1. On the scaling with an autonomously controlled expander
We consider a three-dimensional domain of volume V = AL, in which the graded
morphogen profile forms along the x-direction of length L, whereas A denotes the
cross-sectional area. Now, we aim to achieve scaling with respect to some part of the
system L1 with L = L1 + L2. For example, L1 could be the target tissue without the
morphogen source (if L2 = w) or the whole system (if L2 = 0).
The expander dynamics are described by
∂tE = DE ∇2 E − βE E + νE
(A.11)
with a constant diffusion term DE and the terms for degradation βE = βE(x) and
production νE = νE(x) that depend on the position in the tissue. In the limit of a fast
diffusing expander, the general steady state of Eq. A.11 is given by
where
(cid:90)
(cid:104)βE(cid:105) =
1
V
E∗ = (cid:104)νE(cid:105) /(cid:104)βE(cid:105) ,
βE dV
and (cid:104)νE(cid:105) =
(cid:90)
1
V
(A.12)
sE dV
(A.13)
denote the spatial averages. In order to achieve scaling of the morphogen profile with
respect to L1, the expander level has to be a function of L1 in the steady state. Hence,
(cid:104)νE(cid:105) and (cid:104)βE(cid:105) must not depend on L1 in the same way.
104
A.2 Gradient scaling with expander
If the expander is degraded everywhere in the system part of length L1 und produced
in a source of constant width wE like in Eq. 2.9, we recover the result from Eq. 2.10 by
computing
(cid:104)βE(cid:105) =
βE A L1
,
(cid:104)νE(cid:105) =
V
αE A wE
V
.
(A.14)
The steady state of the expander would be a function of L1 and thus could couple to the
morphogen dynamics to generate scaling. Analogously, degradation could happen only
at the boundary in a stripe of width wE and production might be turned on everywhere
along the length L1, thus
(cid:104)βE(cid:105) =
βE A wE
V
would also make E∗ a function of L1.
(cid:104)νE(cid:105) =
,
αE A L1
V
(A.15)
A.2.2. Scaling by expander feedback with a switch-like production
We assume that the expander is suppressed by the morphogen in a switch-like manner
according to Eq. 2.9 and discuss implications for scaling and robustness of the feedback
schemes.
By combining Eq. 2.14 and the steady state solution for reflecting boundary conditions
given by Eq. 2.3, we obtain
sinh(w/λ∗)
sinh(L/λ∗)
α
β∗
cosh(wE/λ∗) = Cth .
(A.16)
We can derive a relationship between β∗ and E∗ by replacing w∗
E using Eq. 2.15
β∗ =
α
Cth
sinh(w/λ∗)
sinh(L/λ∗)
cosh
(cid:18) βE L E∗
(cid:19)
αE λ∗
.
(A.17)
In order to illustrate the behavior of the feedback system, we are considering the two
limiting cases of (i) a scaling morphogen source with w = χwL and (ii) a constant
morphogen source. For both, we assume that there exists a perfectly scaling steady
state with λ∗ = χλL and a constant scaling factor χλ. Next, we show what this implies
for the coupling between expander and morphogen and discuss the consequences.
(i) For a scaling morphogen source, the following relation between β∗ and E∗ has to
hold in order for the morphogen profile to scale in the steady state:
β∗ =
α
Cth
sinh(χw/χλ)
sinh(1/χλ)
cosh
.
(A.18)
(cid:18) βE
αEχλ
E∗(cid:19)
The coefficient χλ is a free parameter encoding the effect of the expander on β. Note
that it is not possible to achieve scaling by only adjusting the diffusion coefficient D
105
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
instead of β because this equation does not depend on D. Eq. A.18 challenges the
stability of the feedback loop because the expander has a suppressing effect on the
morphogen (at least in the steady state), while according to our Eq. 2.9 the morphogen
also suppresses the expander. Thus, this feedback tends to be unstable.
(ii) For a constant morphogen source, we obtain
(cid:32)
sinh(w(cid:112)β/D)
β
(cid:33)∗
=
α
Cth sinh(1/χλ)
cosh
.
(A.19)
(cid:18) βE
αEχλ
E∗(cid:19)
Now, either of both, β or D, can be adjusted for scaling. The question on the stability
is a bit more complicated than in the case above, where an increase in β results in a
decrease of the morphogen level everywhere in the system. In contrast, for example,
an increase in D results in a decrease of the morphogen level close to the morphogen
source and an increase at the other side of the system because the morphogen is dis-
tributed more homogeneously. In consequence, the feedback can be stable in a limited
size range, as long as the resulting expander source is sufficiently far away from the
morphogen source. As the size of the expander source changes stronger than linearly
with L (compare to Eq. 2.10), the system will eventually reach a size at which this is
not fulfilled anymore. Analogously, if β is adjusted, there also is a size limit. At this
size, β is too small such that the effect of E on β is positive according to Eq. A.19.
The size limit corresponds to λ larger than w, which is likely the case for most of the
systems as w is kept constant.
A.2.3. Scaling by expander feedback with a graded production
The step-like production term in Eq. 2.12 yields rather complicated relationships be-
tween expander and morphogen dynamics, see Eq. A.18 and A.19.
Instead, when
considering a non-linear production term
∂tE(t, x) = DE ∂2
x E(t, x) − βE E(t, x) + αE C(x)h
(A.20)
with an arbitrary real number h, one obtains analogous results in a more general and
much clearer fashion. The exponent h can be positive or negative, corresponding to
an enhancing or a repressing effect on the source size, respectively. The term can
be understood as an approximation to a Hill function for concentrations below the
saturation.
The expander concentration in the steady state for the limit of fast diffusion is
(cid:90) L/λ∗
E∗ =
αE λ∗
βE L
0
106
C∗(x(cid:48))h dx(cid:48) with x(cid:48) = x/λ∗ .
(A.21)
A.3 Linear stability analysis of a Turing system
Again, we will distinguish between the limiting cases of (i) a scaling morphogen source
and (ii) a morphogen source of constant size.
(i) For a scaling source, we obtain from Eq. 2.3
(cid:18) α
β∗
(cid:19)h I1(L/λ∗) ⇒ β∗ ∝ E∗ −1/h ,
E∗ =
αE
βE
(A.22)
where I1 is a short-hand form of the integral term. In analogy to the switch-like source
in Section A.2.2, we cannot obtain scaling by solely adjusting the diffusion properties.
If the degradation rate is adjusted, the steady state tends to be unstable, irrespective
of the exponent h, which characterizes whether the morphogen has an enhancing or
suppressing effect.
(ii) If we assume w is constant, it follows
(cid:32)
α sinh(w(cid:112)β/D)
(cid:33)∗ h
I2(L/λ∗) + O(w/λ∗)2
E∗ =
αE
βE
β
⇒ (β D)∗ ∝ E∗ −2/h
for w (cid:28) λ∗
(A.23)
(A.24)
with I2 denoting the integral. A corresponding expression like in Eq. A.24 can be
derived without the limit w (cid:28) λ if there is no morphogen degradation in the source,
see Appendix A.1.4. As before, there can be a stable scaling fixed point if either β or
D is controlled, yet only in a limited size range.
A.3. Linear stability analysis of a Turing system
A.3.1. Eigenvalues of the linearized reaction-diffusion matrix
The linearized reaction-diffusion system in Fourier space given by Eq. 1.2 is character-
ized by the 2 × 2-matrix Ms in Eq. 1.4:
(cid:33)
(cid:32)M 1
s M 2
s
M 3
s M 4
s
Ms =
(A.25)
The two eigenvalues qI
s of Ms determine the linear stability of the homogeneous
steady state, about which the system is linearized. If the real parts of both eigenvalues
s , qII
are negative, the steady state is stable, otherwise the steady state is unstable.
107
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
Figure A.1.: (a) The trace-determinant plot summarizes the properties of the two
eigenvalues of the 2 × 2-matrix Ms. (b) The two eigenvalues of the matrix Ms for the
parameters used in Section 2.5 with finite DB (red) and DB → ∞ (blue). For our
choice of parameters and reflecting boundary conditions only the first mode (s = 0.5)
is linearly unstable, compare to Fig. 2.8. Parameters: DB/DA = 30, αB/αA = 4,
1.5 ≈ 1.2, h = 5 (if not stated otherwise).
βB/βA = 2, λA/L =
0.1 ≈ 0.3, λB/L =
√
√
The eigenvalues are given by
(cid:16)
(cid:16)
s − qI/II
s
M 1
M 4
(cid:17)2 − Tr[Ms] qI/II
qI/II
s
(cid:17)(cid:16)
s
s M 2
s = 0
s − qI/II
(cid:17) − M 3
(cid:115)(cid:18) Tr[Ms]
(cid:19)2 − Det[Ms] ,
s + Det[Ms] = 0
2
qI/II
s
=
Tr[Ms]
2
±
(A.26)
where "Tr" denotes the trace and "Det" denotes the determinant. The properties of
the eigenvalues can be summarized in the Tr-Det-diagram in Fig. A.1(a). If Tr[Ms]2 <
4 Det[Ms], both eigenvalues are complex and, in fact, are complex conjugates of each
other. For Tr[Ms] < 0 both real parts are negative, otherwise both real parts are
positive.
Instead, if Tr[Ms]2 < 4 Det[Ms], both eigenvalues are real. For Det[Ms] > 0 both
eigenvalues have the same sign: a positive sign for Tr[Ms] > 0 and a negative sign for
Tr[Ms] < 0. This can be easily seen from
qI
s + qII
s = Tr(Ms)
and qI
s · qII
s = Det(Ms) .
(A.27)
108
both eigenvalues complex with positive real partsboth eigenvalues real and negativeboth eigenvalues real and positiveboth eigenvalues real but of opposite signboth eigenvalues real but of opposite signstablestableunstableunstableunstableunstableTraceDet(a)(b)both eigenvalues complex with negative real parts1432-0.50.5-115-5-10A.3 Linear stability analysis of a Turing system
Finally, for Det[Ms] < 0 both eigenvalues are real but of opposite sign.
For the homogeneous steady state to be stable with respect to homogeneous perturba-
tion (s = 0), both real parts of the eigenvalues qI
0 have to be negative. Thus, the
0, qII
matrix M0 should be located in the second quadrant, where
Tr(M0) < 0
and Det(M0) > 0 .
(A.28)
For the homogeneous steady state to be unstable with respect to inhomogeneous per-
turbation (s (cid:54)= 0), at least one of the eigenvalues qI
s has to have a positive real
part. It turns out that by changing s, the trace cannot become positive if Tr(M0) < 0:
s , qII
s + M 4
Tr(Ms) = M 1
(A.29)
Thus, the instability for s (cid:54)= 0 has to correspond to the third quadrant, for which both
eigenvalues are real and one of them is positive:
s = M 1
0 + M 4
0 − (DA + DB)(2πs/L)2 < 0 .
Det(Ms) < 0 with s (cid:54)= 0 .
(A.30)
Note that Tr(Ms) decreases with increasing s, while Det(Ms) first decreases and later
increases again according to
Det(Ms) = Det(M0) − (DA M 4
0 + DB M 1
0 )(2πs/L)2 + DA DB (2πs/L)4 .
(A.31)
As a consequence, the eigenvalues of a Turing system depend on the mode number s as
depicted in Fig. A.1(b). For an intermediate range of s one eigenvalue is positive and
the system is linearly unstable with respect to the corresponding modes.
A.3.2. The principle of local activation and lateral inhibition
We derive necessary conditions for spontaneneous pattern formation in a two-component
reaction diffusion system. This section is based on analogous derivations in the litera-
ture (64, 65, 146, 195, 221). For the following, we define
(cid:12)(cid:12)A∗
(cid:12)(cid:12)A∗
h
h,B∗
h,B∗
h
RAA = ∂ARA
RBA = ∂ARB
, RAB = ∂BRA
, RBB = ∂BRB
(cid:12)(cid:12)A∗
(cid:12)(cid:12)A∗
,
.
h
h,B∗
h,B∗
h
(A.32)
(A.33)
(A.34)
The conditions in Eq. A.28 and A.30 result in the relations
(cid:16)
RAA − DA(2πs/L)2(cid:17)(cid:16)
RAA + RBB < 0
RAA RBB − RAB RBA > 0
RBB − DB(2πs/L)2(cid:17)
< RBA RAB .
(A.35)
109
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
The last two inequalities can be combined to
RAA DB > −RBB DA + DA DB (2πs/L)2 .
(A.36)
Eq. A.33 tells us that at least one of the two species has to have a self-inhibiting effect.
Here, we choose species B: RBB < 0. From Eq. A.36, we see that the other one, i.e. A,
has to be self-enhancing: RAA > 0. From Eq. A.34 follows that RAB RBA < 0. The
cross-reaction terms must be of opposing signs.
As RBA RAB < 0 and RBB − DB(2πs/L)2 < 0 in Eq. A.35, it follows that
RAA − DA(2πs/L)2 > 0 .
(A.37)
The diffusion coefficient DA of the self-activator has to be sufficiently small for certain
modes to become unstable ("local activation"). This becomes even more apparent if we
consider the reaction term RA = αAP (A, B) − βA A like in Eq. 2.18. Now, the char-
acteristic wavelength of the reaction-diffusion equation has to be sufficiently small in
/βA − 1)/(2πs)2. The weaker
the self-activating effect and the larger the wave number s of the perturbation, the
comparison to the system size: (λA/L)2 < (αA∂AP(cid:12)(cid:12)A∗
h,B∗
h
smaller λA has to be.
Finally, from Eq. A.36, we obtain RAA DB > −RBB DA. With Eq. A.33, hence
RAA < −RBB, we see that the self-inhibitor has to diffuse faster than the self-activator
("lateral inhibition"):
DB > DA .
(A.38)
A.4. Motivation for the Hill function
The production function P (A, B) is a switch-like element that takes two inputs and com-
pares them. In a biological system the inputs are the concentrations of two molecules
that bind to the receptors of a cell. We discuss two possibilities to compare the concen-
trations of activator A and the inhibitor B by simple generic mechanisms that describe
binding and unbinding to receptors. The number of unbound receptors on a cell is
given by Q. The output could be the number of receptors QA that are bound to the
activator molecule.
1. Case: The activator A binds to the receptor Q while the inhibitor B leads to
unbinding of A. The dynamic equation for concentration of receptors bound to A is
QA = γ1 Ah1 Q − γ2 Bh2 QA − γ3 QA ,
(A.39)
110
A.4 Motivation for the Hill function
where h1 and h2 describe cooperativity effects for binding and unbinding. For example,
if the activator only binds in pairs, h1 = 2. The coefficients γ1, γ2 and γ3 are the rates
of binding as well as induced and spontaneous unbinding. If the total concentration of
receptors Q0 = Q + QA is constant, the steady state reads
QA =
γ1 Ah1 Q0
γ3 + γ2 Bh2 + γ1 Ah1
h=h1=h2
−−−−−−→
γ3=0
γ1 Q0
(A/B)h
γ2/γ1 + (A/B)h .
(A.40)
If the cooperativity exponents are the same and the activator rarely unbinds spona-
neously, the standard Hill function is recovered. A finite γ3 (cid:54)= 0 allows to compute the
limit A = B = 0.
2. Case: We have seen that if the inhibitor facilitates the unbinding of the activator,
we obtain a Hill function for the concentration of receptors bound to the activator.
Another option for implementing the inhibiting effect is competitive binding. Here, the
inhibitor binds to the same receptors as the activator and thus blocks activator binding.
Now, the dynamic equations are
QA = γ1 Ah1 Q − γ2 QA
QB = γ3 Bh2 Q − γ4 QB
Q = −(γ1 Ah1 + γ3 Bh2) Q + γ2 QA + γ4 QB .
In the steady state Eq. A.41 results in
Q =
γ2
γ1 Ah1
QA
(A.41)
(A.42)
(A.43)
(A.44)
and Eq. A.42 yields
γ3 Bh2(Q0 − QA)
γ3 Bh2
γ4
Q =
QB =
(A.45)
with Q = Q0 − QA − QB, again assuming a constant total receptor number Q0. When
inserting these two equations in the steady state of Eq. A.43, we finally get
γ3 Bh2(Q0 − QA)
γ4 + γ3 Bh2
−(γ1 Ah1 + γ3 Bh2)
γ2
γ1 Ah1
QA + γ2 QA + γ4
γ4 + γ3 Bh2
= 0
γ2 γ3 Bh2
γ1 Ah1
QA + γ4
γ3 Bh2
γ4 + γ3 Bh2
QA = γ4
γ3 Bh2
γ4 + γ3 Bh2
Q0
γ2 γ4 + γ2 γ3 Bh2
γ4 γ1 Ah1
QA + QA = Q0
QA =
γ4 γ1 Ah1 Q0
γ2 γ4 + γ2 γ3 Bh2 + γ4 γ1 Ah1
.
111
(A.46)
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
This is also a Hill function with exactly the same structure as the one above. It becomes
apparent that this function stands for a wide class of realizations of a logical element
that compares the magnitude of two inputs in a biological system.
A.5. Homogeneous steady state of our Turing system
The homogeneous steady state of our choice of Turing system given by Eq. 2.18 is
characterized by
B∗
h
(A.47)
A∗
h =
αAβB
αBβA
h, B∗
βA
αA
Fig. A.2 illustrates the steady state in terms of A∗
h for the Hill function of Eq. 2.19 and
the step function of Eq. 2.20. The solid red line corresponds to the left side and the
(A.48)
h) =
P (A∗
A∗
h .
dotted line to the right side of Eq. A.48. The intersections mark the steady state value
of A∗
We use a specific definition of the theta function at position zero
h. There is a second case for the step function (purple dashed line).
0
Θ(0) !=
if αAβB
αBβA
if αAβB
1
αBβA
h ≤ B∗
h and A∗
> 1 .
h > B∗
≤ 1
(A.49)
h,B∗
h
=
h,B∗
h
h βA/αA
h/B∗
h)h
1 + (A∗
∂AP(cid:12)(cid:12)A∗
= − A∗
B∗
h
h
h,B∗
h
(A.51)
(A.52)
.
112
These two cases correspond to A∗
h, respectively. This definition is
chosen to ensure the existence of a steady state and to avoid artificial oscillations in the
simulations. Moreover, it establishes the correspondence between the theta function
and the Hill function in the limit of h → ∞.
Next, we will show that only the case of A∗
h ensures spontaneous pattern formation
irrespectively of h. For this, we explicitly perform a linear stability analysis of our
h ≤ B∗
specific Turing system. The matrix of Eq. A.25 for the linearized system reads
αA ∂AP(cid:12)(cid:12)A∗
Ms =
− βA − DA(2πs/L)2
αA∂BP(cid:12)(cid:12)A∗
h,B∗
h
αB∂BP(cid:12)(cid:12)A∗
h,Bh
− βB − DB(2πs/L)2
.
(A.50)
It includes the derivatives of the Hill functions with respect to A and B at steady state:
h,B∗
h
h,B∗
h
αB ∂AP(cid:12)(cid:12)A∗
∂AP(cid:12)(cid:12)A∗
∂BP(cid:12)(cid:12)A∗
A.5 Homogeneous steady state of our Turing system
Figure A.2.: Homogeneous steady state of our Turing system for (a) a Hill function
and (b) a theta function. Solid red (and dashed purple) lines correspond to the left
side of Eq. A.48, dotted red lines to the right side.
The limit of h → ∞ diverges for αAβB ≤ αBβA:
∂AP(cid:12)(cid:12)A∗
→
h,B∗
h
∞ if αAβB
αBβA
if αAβB
αBβA
0
≤ 1
> 1
.
(A.53)
This is in in agreement with the theta function, for which the derivative diverges at
position zero and is zero everywhere else.
The determinant of the matrix Ms is
(cid:16)
βA + DA(2πs/L)2(cid:17)(cid:16)
βB + DB(2πs/L)2(cid:17) − βA βb (2πs/L)2 h
(cid:16)
(cid:17)
.
B − λ2
λ2
A
Det(Ms) =
1 + (A∗
h/B∗
h)h
(cid:16)
B − λ2
λ2
A
(cid:17)
(A.54)
.
(A.55)
Thus, the instability condition of Eq. A.30 is given by
(cid:16)
βA + DA(2πs/L)2(cid:17)(cid:16)
βB + DB(2πs/L)2(cid:17)
<
βA βb (2πs/L)2 h
1 + (A∗
h/B∗
h)h
First of all, this requires λB > λA, typical for Turing systems. Furthermore, in order
to fulfill this inequality for arbitrary h and in particular in the limit h → ∞, it follows
that
h ≤ B∗
A∗
h ,
αAβB
αBβA
≤ 1
(A.56)
Our choice of parameters in Chapter 2 amounts to (αAβB)/(αBβA) = 1/2. The second
set of conditions of Eq. A.28 result in
h
1 + (A∗
h/B∗
h)h (βA − βB) < βA + βB
βA + βB > 0
(A.57)
(A.58)
113
(a)(b)1A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
While Eq. A.58 is trivially fulfilled, Eq. A.57 requires βA < βB in order to hold for
arbitrarily large h. In our parameter sets of Chapter 2, we have chosen βB/βA = 2.
A.6. Inhomogenous steady states of our Turing system
A.6.1. First order steady state solution
First, we solve Eq. 2.18 of our Turing system with a step-like production function by
assuming a single source region which is touching the boundary as shown in Fig. A.3(a).
We refer to this as the (1, 1)-pattern. The general steady state solution for A∗
(1,1) is a
piecewise function
A∗
in(x) = Ain,1 ex/λA + Ain,2 e−x/λA +
A∗
out(x) = Aout,1 ex/λA + Aout,2 e−x/λA
αA
βA
for 0 < x ≤ (cid:96)
(A.59)
for (cid:96) < x < L
(A.60)
for the region inside and outside the source. In order to determine the four constant
factors, we use the two reflecting boundary conditions and two continuity conditions at
the source boundary:
out(x)(cid:12)(cid:12)(cid:96) .
out((cid:96))
∂xA∗
∂xA∗
A∗
in((cid:96)) = A∗
∂xA∗
in(x)(cid:12)(cid:12)0 = 0
out(x)(cid:12)(cid:12)L = 0
in(x)(cid:12)(cid:12)(cid:96) = ∂xA∗
(cid:18) x
(cid:19)
(cid:18) x − L
λA
+
λA
(A.61)
(A.62)
(A.63)
(A.64)
(A.65)
(A.66)
(A.67)
From Eq. A.61-A.62, we obtain Ain,1 = Ain,2 and Aout,1 = Aout,2 e−2L/λA, respectively.
Thus, the solution becomes
A∗
in(x) = Ain,0 cosh
A∗
out(x) = Aout,0 cosh
αA
βA
(cid:19)
for 0 ≤ x ≤ (cid:96)
for (cid:96) < x ≤ L
1 − sinh(L/λA−(cid:96)/λA)
cosh(x/λA)
sinh(L/λA) cosh(x/λA − L/λA)
sinh(L/λA)
sinh((cid:96)/λA)
for 0 ≤ x ≤ (cid:96)
for (cid:96) < x ≤ L .
A∗
(1,1) =
αA
βA
Finally, from Eq. A.63-A.64, we compute the (1, 1)-solution
The solution for B∗
(1,1) is determined analogously, see Eq. 2.24.
114
A.6 Inhomogenous steady states of our Turing system
Figure A.3.: (a) The concentration profile of A∗
(1,1) can be written as a piecewise
function for the part inside and outside the source. (b) The intersections (circles) of
the blue and the red curves mark the steady state source size according to Eq. A.69.
√
Parameters like in Section 2.5: DB/DA = 30, αB/αA = 4, βB/βA = 2, λA/L =
0.1 ≈ 0.3, λB/L =
1.5 ≈ 1.2.
√
A.6.2. Source size of the first order steady state
For the step-like production function, the source size is defined by
A∗
(1,1)((cid:96)) = B∗
FA((cid:96)) =
(1,1)((cid:96))
with
F((cid:96)) =
sinh((cid:96)/λ)
sinh(L/λ)
cosh((cid:96)/λ − L/λ) =
1
2
αBβA
αAβB
FB((cid:96))
(cid:18)
1 − sinh(L/λ − 2(cid:96)/λ)
sinh(L/λ)
(A.68)
(A.69)
(cid:19)
.
(A.70)
Next, we show that the source size (cid:96) has only one non-trivial solution. First, we compute
F(0) = 0 , F(L/2) = 1/2 , F(L) = 1 ,
(A.71)
and we see that (cid:96) = 0 is always a solution to Eq. A.69. The curve F((cid:96)) is essentially a
shifted hyperbolic sine function und is monotonously increasing
∂(cid:96)F((cid:96)) =
cosh(2(cid:96)/λ − L/λ)
λ sinh(L/λ)
> 0 .
The largest slope is found at the interval bounds:
∂(cid:96)F((cid:96))(cid:12)(cid:12)0 = ∂(cid:96)F((cid:96))(cid:12)(cid:12)L =
1
λ tanh(L/λ)
.
(A.72)
(A.73)
115
(a)(b)0010.5Source sizeA. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
In particular, we see that
(cid:18)
(cid:19)
∂λ
1
λ tanh(L/λ)
2 L/λ − sinh(2L/λ)
2 λ2 sinh(L/λ)2 < 0 because L/λ > 0 .
=
(A.74)
Thus, the larger λ the smaller the slope at (cid:96) = 0 and (cid:96) = L.
In consequence, the
curve for λB lies below the curve for λA if (cid:96) < L/2 and above if (cid:96) > L/2, as shown in
Fig. A.3(b). We can demonstrate this fact explicitly by computing
(cid:18) sinh(2(cid:96)/λ)
(L − (cid:96)) (cid:96)
∂λF((cid:96)) =
λ3 sinh(L/λ)2
2(cid:96)/λ
(cid:19)
− sinh(2L/λ − 2(cid:96)/λ)
2(L − (cid:96))/λ
≶ 0
if (cid:96) ≶ L/2 .
(A.75)
Therefore, FA((cid:96)) and FB((cid:96)) always have three intersection points at 0, L/2 and L. If
we now consider Eq. A.69 with αAβB < αBβA and λA < λB (as required for the Turing
system to form patterns, see Section A.5), (cid:96) = L is not a solution anymore and the
second solution is shifted to smaller values of (cid:96) (intersection with dashed blue line in
Fig. A.3(b)).
In summary, there is always a solution with (cid:96) = 0 (no source). Additionally, for our
Turing condition αAβB ≤ αBβA, there can be a second solution with a finite source
size (cid:96) < L/2 if
αBβA
αAβB
λA tanh(L/λA) < λB tanh(L/λB) .
(A.76)
For αAβB < αBβA, there are no further solutions. The condition of Eq. A.76 marks a
bifurcation point of the system, beyond which the (1, 1)-solution exists in addition to
the (0, 0)-solution.
From Eq. A.76, we see that λA is constrained by either L or λB, depending on which
one is smaller. If L (cid:28) λB, the equation becomes
αAβB
αBβA
tanh(L/λA) <
L/λA .
(A.77)
For αAβB < αBβA, this is only fulfilled if L/λA is large enough. In contrast for L (cid:29) λB,
we obtain
λA <
λB .
(A.78)
(cid:18)
For αAβB < αBβA, this is only fulfilled if λA is sufficiently smaller than λB.
Finally, we consider two limiting cases of Eq. A.70:
F((cid:96)) =
1
2
1 − sinh(L/λ (1 − 2(cid:96)/L))
sinh(L/λ)
→
(cid:96)/L for L/λ (cid:28) 1
for L/λ (cid:29) 1 .
1/2
(A.79)
Note that the absolute value of 1 − 2(cid:96)/L is smaller than 1. Thus, for L (cid:28) λB and
L (cid:29) λA, we obtain (cid:96)/L = (αAβB)/(2αBβA), see Fig. 2.7.
116
αAβB
αBβA
(cid:19)
(cid:40)
A.6 Inhomogenous steady states of our Turing system
A.6.3. Hierarchy of higher order steady states
We can successively derive the higher order steady states. First, we consider the
second order solution with two sources at the boundaries, denoted as (2, 2)-pattern,
see Fig. A.6.4(a). Thus, there are maxima of concentration A∗
(2,2) at each
boundary. Furthermore, there is a minimum for each concentration in between the two
(2,2) and B∗
source regions. We now prove that the system is completely symmetric as depicted in
Fig. A.6.4(b). For that it is sufficient to show that the minima for A∗
(2,2) are
at the same position, i.e. LA = LB. Let us consider the solutions left of the respective
(2,2) and B∗
minima in Fig. A.6.4(a). These solutions correspond to the (1, 1)-pattern, which we
have characterized above. If the minima are at the same position (LA = LB), there is
at most one (cid:96) with 0 < (cid:96) < L in agreement with A∗
(1,1)((cid:96)), see Section A.6.2.
However, B∗
(1,1)((cid:96)) is a strictly monotonic function of LB if (cid:96) > 0:
(1,1)((cid:96)) = B∗
∂LB B∗
(1,1)((cid:96)) = − αB sinh((cid:96)/λB) cosh((cid:96)/λB)
sinh(L/λB)2λB
< 0 .
(A.80)
Therefore, there is no other choice of LB also solving the equation for (cid:96). In consequence
we conclude that LA = LB in the steady state.
For LA = LB, we can cut the (2,2)-pattern at this position, which results in two (1,1)-
patterns. But since these two (1,1)-patterns have in fact the same concentration A at
their common boundary, their system size must be equal. An analogous reasoning can
be applied to the (1, 0)-pattern. Finally, Fig. A.6.4(b) exemplifies that any higher order
pattern can be considered as a concatenation of (1, 1)-patterns as the basic building
blocks. Each adjacent pair of these fundamental sections is either a (1, 0)-pattern or a
(2, 2)-pattern and thus completely symmetric. In consequence, any higher order pattern
can be constructed by a concatenation of (1, 1)-patterns of identical size.
A.6.4. Stability of the inhomogeneous steady state of our Turing system
We are probing the stability of the inhomogeneous steady states in various ways. First,
we can add a set of small perturbations to the steady state profiles and monitor the
relaxation behavior in numerical simulations. Second, we can start with a linear com-
bination of two steady state patterns with varying weights and observe to which steady
state they converge, as shown in Fig. 2.8(b). Finally, for the case of the Hill-type pro-
duction function, we can numerically perform a linear stability analysis, as we will show
next.
117
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
Figure A.4.: (a) The concentrations A and B of the (2, 2)-pattern possess maxima
at boundaries and one minimum each between the source regions. (b) All patterns
can be constructed by a concatenation of (1, 1)-patterns of identical size. Parameters
0.1 ≈ 0.3,
like in Section 2.5: DB/DA = 30, αB/αA = 4, βB/βA = 2, λA/L =
λB/L =
1.5 ≈ 1.2.
√
√
Let us consider small perturbations a(t, x) and b(t, x) about the inhomogeneous steady
states:
A(t, x) = A∗
(m,n)(x) + a(t, x) , B(t, x) = B∗
(m,n)(x) + b(t, x) .
The linearized dynamics of the perturbations is given by
with the operators
(cid:32)P1 P2
(cid:33)(cid:32)a
(cid:33)
P3 P4
b
(m,n),B∗
(m,n),B∗
(m,n),B∗
(m,n),B∗
(m,n)
(m,n)
(m,n)
(m,n)
− βA + DA∂2
x
− βB + DB∂2
x .
b
=
∂t
(cid:33)
(cid:32)a
P1 = αA ∂AP(cid:12)(cid:12)A∗
P2 = αA∂BP(cid:12)(cid:12)A∗
P3 = αB ∂AP(cid:12)(cid:12)A∗
P4 = αB∂BP(cid:12)(cid:12)A∗
(cid:88)
(A.81)
(A.82)
(A.83)
(A.84)
(A.85)
(A.86)
Now, we express the perturbations in terms of a set orthonormal modes that agree with
the reflecting boundary conditions:
a(t, x) = a0 +
aj(t) mj(x) with mj(x) =
√
2 cos(π j x/L) .
(A.87)
j=1
118
(a)(b)A.6 Inhomogenous steady states of our Turing system
Figure A.5.: Maximum eigenvalue of the linear operator matrix for the (1, 0)-pattern
as a function of system size. The pattern becomes stable at L/λA ≈ 5.5, see also
Fig. 2.8. Approximation with 20 modes and a spatial discretization of the profiles with
200 grid points. Parameters like in Fig. 2.8: DB/DA = 30, αB/αA = 4, βB/βA = 2,
h = 5.
The coefficients aj(t) can be determined by
(cid:90) L
0
aj(t) =
a(t, x) mj(x) dx .
(A.88)
The linearized system in terms of these modes is
=
(cid:82) L
(cid:82) L
a1
a2
...
b1
b2
...
∂t
(cid:82) L
(cid:82) L
0 mjP1mi dx
0 mjP2mi dx
0 mjP3mi dx
0 mjP4mi dx
a1
a2
...
b1
b2
...
(A.89)
If the largest eigenvalue of the linear operator matrix is negative, the steady state
is stable and this maximum eigenvalue provides the time scale of relaxation of the
slowest decaying mode. If at least one eigenvalue is positive, the steady state is linearly
unstable.
Fig. A.5 shows the maximum eigenvalue for the (1, 0)-pattern as a function of the system
size. We can determine the lower bound for which this pattern becomes stable, as
illustrated in Fig. 2.8. For none of the inhomogeneous patterns, we observe a maximum
system size, at which these patterns would become unstable again.
119
4.555.566.5-0.0600.060.12System sizeMaximum eigenvalueA. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
A.7. On our scalable Turing system
A.7.1. A homogeneous dynamic state for low expander levels
For low expander values, the concentrations in the Turing system with expander feed-
back become homogenous. Here, we consider the adiabatic limit for which the expander
relaxation is much faster than the dynamics of the Turing molecules. In consequence,
the expander tightly follows the much slower dynamics of the inhibitor concentration
B:
E =
αE
κE B
.
(A.90)
It exists a dynamic state of the Turing system for which the concentrations A and
B relax in synchrony such that the ratio χ = B/A stays constant. As a result, the
Hill-type production function has a constant value P (A, B) = g(χ) given by
g(χ) =
1
1 + χh .
The ratio χ is is obtained from
∂tχ = 0
αBg(χ) − κBEB − αAg(χ)χ + κAEB = 0
g(χ) =
κA − κB
αA χ − αB
.
αE
κE
(A.91)
(A.92)
The system can leave the homogeneous regime if E increases. For this, B has to
decrease, according to Eq. A.90. We have discussed for Eq. 3.11 that this requires
g(χ) < (cid:96)∗/L, where (cid:96)∗/L = (αEκB)/(κEαB). For the threshold where g(χ) = (cid:96)∗/L, we
obtain
κA − κB
αA χ − αB
αE
κE
=
αEκB
κEαB
χ =
αBκA
αBκB
.
(A.93)
Thus, the ratio B/A equals the ratio of the homogeneous steady state concentrations
B∗
h/A∗
h. Therefore, from g(χ) = (cid:96)∗/L follows that f(0,0) = (cid:96)∗/L.
120
A.7 On our scalable Turing system
A.7.2. Generalized scalable Turing system
We can generalize our approach of scalable Turing patterning by considering the fol-
lowing generic equations for A, B and E
∂tA = αA P (A, B) + RA(A, E) + DA ∂2
∂tB = αB P (A, B) + RB(B, E) + DB ∂2
x A
∂tE = αE + LE
RA(A, E), RB(B, E)
(cid:16)
(cid:17)
x B
+ DE ∂2
x E .
(A.94)
(A.95)
(A.96)
Here, RA and RB are generic functions fulfilling the Turing conditions and LE is a
linear function of RA(A, E) and RB(B, E). By computing the spatial averages in the
steady state, one obtains
αE + (cid:104)LE
αE + LE
RA(A, E), RB(B, E)
(cid:17)(cid:105) = 0
(cid:16)
(cid:16)(cid:104)RA(A, E)(cid:105),(cid:104)RB(B, E)(cid:105)(cid:17)
(cid:16)
αA(cid:104)P(cid:105), αB(cid:104)P(cid:105)(cid:17)
= 0 .
= 0
αE + LE
(A.97)
If RA, RB and LE are chosen such that LE((cid:104)P(cid:105)) = LE(αA(cid:104)P(cid:105), αB(cid:104)P(cid:105)) is a monotonic
function in (cid:104)P(cid:105), this uniquely determines the source size in the steady state. Thus, if
a steady state exists, the size of the source will be independent of the system size.
Let us again consider a homogeneous expander distribution. Furthermore, we assume
that the relaxation of the Turing system is much faster than the relaxation of the
expander. Thus, Eq. A.94-A.95 define the size of the source as a function of slowly
varying E. The intersection of this nullcline with the solution for (cid:104)P(cid:105) from Eq. A.97
provides the steady state values of E. For the steady state to be linearly stable, it
requires that ∂(cid:104)P(cid:105) LE((cid:104)P(cid:105)) · ∂E(cid:104)P(cid:105)E∗ < 0.
A.7.3. Scaling of downstream targets with a constant amplitude
One important feature of our scalable Turing system is that the amplitude of the
morphogens increases quadratically with L, see Section 3.7. This is a general property
of such scaling mechanisms for which the degradation rate is adjusted, see Section 2.4.5.
As the degradation rate changes proportional to L−2, the products βAA and βBB are
characterized by a constant amplitude independent of system size. If the cell responds
to this flux, the expression of down-stream targets scales with a constant amplitude.
Yet, how might the cell read out the degradation flux?
Signaling often requires binding of the molecules to receptors on the cell surface. The
concentration of activator molecules bound to receptors is Ab ∝ βAb A, where βAb
121
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
is the binding rate. If the molecules become internalized or degraded upon binding,
this binding removes the molecules from the system and the binding rate βAb in fact
contributes to the degradation rate βA.
As a simple example, let us consider the following dynamics of the bound molecules
∂tAb = βAb A − βAr Ab .
(A.98)
Free molecules bind with rate βAb and bound molecules are removed with rate βAr.
The steady state of this equation is
A∗
b =
A∗ .
βAb
βAr
(A.99)
If spontaneous degradation of the free molecules can be neglected in comparison to
binding and internalization, it follows that βA = βAb and the cells read out the flux
βA A∗. In this scenario, the expander could be a co-receptor that controls the binding
rate and thus the effective degradation rate in a size-dependent manner. The down-
stream targets would perfectly scale with a constant amplitude.
A.7.4. On knockout experiments in scalable Turing systems
In Chapter 3, we presented a simple mechanism that couples a classical Turing system
including two molecules of concentration A and B to the dynamics of an additional
expander molecule. Here, we will show that knockout of A or B (for example by RNAi
feedings) can yield misleading results.
If the production of the activator is blocked, one observes that both concentrations fade
away, see Fig. A.6(a). Thus, one would correctly conclude that A activates B. Yet,
after removing the inhibitor from the system, both concentrations decrease as well. In
particular, A is vanishing after an initial increase because of the indirect effect via E.
In consequence one would conclude that B is activating A, despite its direct inhibiting
effect within the Turing feedback loop.
In the case of the flatworms, the Wnt/β-catenin system has an instructive role for tail
identity. RNAi experiments indicate that respective pathway components are expressed
in the tail and have a positive effect on themselves and each other. In contrast, the
inhibitors of this system reside in the head. This is not in agreement with any of the
two most simple Turing systems based on two key players. Fig. 1.10 illustrates that we
would expect either additional inhibiting effects of molecules that are expressed in the
same region or additional activating effects of molecules that are expressed on opposite
ends of worm. Including an expander feedback could solve this problem, as shown in
122
A.7 On our scalable Turing system
Fig. A.6(b). A Turing system of the first type, with both Turing molecules expressed in
the same region, might appear like a mutually enhancing feedback during gene knock-
out. A Turing system of the second type, with both Turing molecules expressed on
opposite ends, might appear like a mutually suppressing feedback.
Monitoring the dynamics of the concentrations upon RNAi feeding over time is more
informative than endpoint assays, see Fig. A.6(a). After knockout of the inhibitor, the
activator first expands before the pattern vanishes, hinting at the suppressive effect of
B.
123
A. REACTION-DIFFUSION SYSTEMS: FIXED POINTS & SCALING
√
√
Figure A.6.: (a) The scalable Turing system discussed in Chapter 3 can appear like
a mutually activating feedback loop in knockout experiments of A and B (dashed lines
are the initial steady states, fading of colors denote the time evolution). Parameters
0.1 ≈ 0.3,
like in Chapter 3: DB/DA = 30, αB/αA = 4, βB/βA = 2, λA/L =
(b) Examples of the two possible Turing topologies
λB/L =
plus an appropriate expander feedback and the interpretation of knockouts of the two
Turing molecules. The results compare to experimental observations with respect to
the Wnt/β-catenin system in flatworms.
1.5 ≈ 1.2, h = 5.
124
(a)ABEPositionConcentration BPositionConcentration AKnockout of A:PositionConcentration BPositionConcentration AABEKnockout of B:(b)seemsto beseemsto beTuring feedback 1:Turing feedback 2:B. On the numerical solution
of reaction-diffusion equations
B.1. Euler method and Courant criterion
A reaction-diffusion equation of type
∂tC(t, x) = −β C(x, t) + D∂2
xC(t, x)
(B.1)
describes diffusion and linear degradation of a chemical species of concentration C in a
one-dimensional domain. It can be solved numerically by discretizing time and space
and using the Euler method as the most simple approach (163). The concentration Ci
at position i at time t + ∆t is then given by
Ci(t + ∆t) = Ci(t) − β Ci(t) ∆t +
D∆t
∆x2
(cid:16)
Ci+1(t) − 2Ci(t) + Ci−1(t)
(cid:17)
.
(B.2)
In order for the Euler update to be numerically stable and not to violate causality, the
following Courant-L´evy-Friedrich criterion has to be fulfilled (163)
∆t <
2∆x2
4D + β∆x2 .
(B.3)
In this thesis, we often analyze systems of several coupled reaction-diffusion equations.
At least one of which usually contains a very fast diffusing chemical species, like the
inhibitor in the Turing system or the expander in Chapter 2 and 3. Usually 4Df (cid:29)
β∆x2 with Df of the fastest diffusing species and β of any species. Thus, ∆t <
∆x2/(2Df ) has to be chosen to be very small. Note that ∆x is bounded from above
especially by the non-homogeneity of the slowest diffusing species. Yet, the time scale
of relaxation to the steady state is determined by the reaction rates β. The number of
iterations to reach the steady state is of order O(cid:2)1/(β∆t)(cid:3) = O(cid:2)2Df /(β∆x2)(cid:3), which
is very large as stated above. Next, we show how to avoid such long computing times.
125
B. ON THE NUMERICAL SOLUTION
OF REACTION-DIFFUSION EQUATIONS
B.2. Algorithmic speed-up using a convolution with a Gauss
kernel
If diffusion of one or several species is fast in comparison to the other processes of the
system, we can apply a separation of time scales to avoid long computing times. During
one rather large time step of the Euler scheme that would violate the Courant criterion
due to the fast diffusion but would be still in agreement with the other slow processes,
we account for the fast diffusion by using the analytical solution of the pure diffusion
equation
A Fourier transform yields
with the solution
∂tC(t, x) = D∂2
xC(x, t) .
∂t C(t, s) = −(2πs)2D C(t, s)
C(t, s) = C0(s) e−(2πs)2Dt
in Fourier space. The back-transform yields an integral with a Gauss kernel
−∞
(cid:90) ∞
(cid:90) ∞
(cid:90) ∞
(cid:90) ∞
−∞
−∞
−∞
C(t, x) =
=
=
=
ds C(t, s) e2iπxs
(cid:90) ∞
−∞
(cid:90) ∞
dy C0(y)
dy C0(y) e
ds e
−∞
−(y−x)2
4Dt
ds
dy C0(y) e−2iπys e−(2πs)2Dt e2iπxs
(cid:16)
(cid:17)2
−4π2Dt
s+ i(y−x)
4π2Dt
−(y−x)2
4Dt
e
1√
4πDt
.
(B.4)
(B.5)
(B.6)
(B.7)
This solution can be used to describe the propagation within one time step of the Euler
scheme.
(cid:90) yj +∆y/2
(cid:88)
yj−∆y/2
j
The solution has to be computed in discrete spatial coordinates. It can be written as
C(t + ∆t, x) =
dy C(t, y) e
−(y−x)2
4D∆t
1√
4πD∆t
.
(B.8)
We define Cj(t) as the average of C(t, y) in the interval [yj − ∆y/2, yj + ∆y/2]. If ∆y
is sufficiently small, such that C(t, y) ≈ Cj(t) (i.e. approximately homogeneous) in the
interval [yj − ∆y/2, yj + ∆y/2], we find
(cid:18)
(cid:20) (yj + ∆y/2) − x
(cid:21)
√
4D∆t
Cj(t)
2
j
Erf
(cid:20) (yj − ∆y/2) − x
(cid:21)(cid:19)
√
4D∆t
− Erf
C(t + ∆t, x) ≈(cid:88)
, (B.9)
126
B.2 Algorithmic speed-up using a convolution with a Gauss kernel
including the error function Erf[z] = 2√
π
concentration at the next time point, we compute the average
(cid:82) z
0 e−ξ2dξ. In order to obtain the discretized
Ci(t + ∆t) =
1
∆x
≈ (cid:88)
j
(cid:90) xi+∆x/2
(cid:32)
xi−∆x/2
Cj(t)
2
−2(i − j) Erf
√
√
4D∆t
π
∆x
+
(cid:18)
dx C(t + ∆t, x)
(cid:21)
(cid:20) (i − j + 1)∆x
(cid:21)
4D∆t
√
+ (i − j − 1) Erf
(i − j + 1) Erf
(cid:20) (i − j)∆x
√
4D∆t
−(i−j+1)2∆x2
4D∆t
e
− 2e
−(i−j)2∆x2
4D∆t + e
(cid:20) (i − j − 1)∆x
(cid:19)(cid:33)
−(i−j−1)2∆x2
4D∆t
√
4D∆t
(B.10)
(cid:21)
.
Here, we used ∆x = ∆y and identified xi−yj = (i−j)∆x. Note that this approximation
is well justified for small ∆x, in particular ∆x <
4D∆t. Next, we demonstrate that
for ∆x (cid:29) √
√
k = i − j and ∆r = ∆x/
Ci(t + ∆t) ≈ (cid:88)
4D∆t the result is not in agreement with Eq. B.4. We can substitute
+ (k − 1) Erf
(k − 1)∆r
(k + 1) Erf
(k + 1)∆r
Ci−k(t)
4D∆t.
(cid:105)
(cid:104)
(cid:104)
(cid:105)
√
e−(k+1)2∆r2 − 2e−k2∆r2 + e−(k−1)2∆r2
(cid:33)
=
Ci(t)
2
2 Erf
∆r
+
(cid:32)
(cid:105)
(cid:104)
k
−2k Erf
k∆r
2
(cid:104)
(cid:32)
(cid:32)
(cid:32)
+
+
+
Ci−1(t)
2
Ci+1(t)
2
(cid:88)
k>2
(k − 1) Erf
+
(cid:105)
(cid:104)
(cid:104)
(cid:32)
√
∆r
π
(cid:33)
(cid:105)
(cid:105)
+
+
∆r
∆r
2 e−∆r2 − 2
√
π
∆r
(cid:105) − 2 Erf
(cid:104)
(cid:105) − 2 Erf
(cid:104)
(cid:104)
2 Erf
2∆r
2 Erf
2∆r
Ci−k(t)
2
(cid:104)
(cid:105)
+
(k − 1)∆r
e−(2∆r)2 − 2 e−∆r2 + 1
∆r
π
√
√
e−(2∆r)2 − 2 e−∆r2 + 1
∆r
(cid:105) − 2k Erf
(cid:104)
π
(cid:105)
(B.11)
(cid:33)
(cid:33)
(cid:33)
(k + 1) Erf
(k + 1)∆r
k∆r
+
e−(k+1)2∆r2 − 2e−k2∆r2 + e−(k−1)2∆r2
√
∆r
π
The Taylor expansion of the error function is given by
Erf[1/x] = sign[x] − e−x2(cid:18) x√
(cid:19)
+ O[x3]
π
.
(B.12)
127
B. ON THE NUMERICAL SOLUTION
OF REACTION-DIFFUSION EQUATIONS
Thus, we obtain
Ci(t + ∆t) ≈ Ci(t) +
= Ci(t) +
π∆r
Ci+1(t) − 2Ci(t) + Ci−1(t)
√
√
D∆t
π∆x
Def f ∆t
√
2
Ci+1(t) − 2Ci(t) + Ci−1(t)
Ci+1(t) − 2Ci(t) + Ci−1(t)
(cid:16)
(cid:16)
(cid:17)
(cid:17)
= Ci(t) +
with Def f = (cid:112)D/(π∆t)∆x. Even though the result looks similar to Eq. B.2 of the
limit ∆x (cid:29) √
Euler method, the diffusion appears to be effectively faster than D. As we consider the
4D∆t, it follows that
(B.13)
∆x2
,
(cid:32)√
(cid:33)2
4D∆t
∆x
<
4D∆t
∆x
√
√
(cid:114)
√
√
D∆t
π∆x
D∆t
2∆x
<
D∆t
∆x2 <
D <
D
π∆t
∆x = DEf f .
(B.14)
The effectively faster diffusion comes from the fact that the concentration within each
interval [yj − ∆y/2, yj + ∆y/2] becomes equally distributed instantaneously by setting
C(t, y) = Cj(t).
In order to avoid this numerical effect, ∆x <
4D∆t is required. In a system with a
slow and a fast diffusion species (Ds (cid:28) Df ), which combines the Euler method with a
Gauss kernel, two conditions have to be fulfilled:
√
∆x2
4Df
< ∆t <
2∆x2
4Ds + βs∆x2 .
(B.15)
128
C. Worm handling and
measurements of size and shape
C.1. Worm handling
If not specified otherwise, we use a clonal line of an asexual strain of Schmidtea mediter-
ranea (28, 185). Worms were maintained at 20◦C as described in (35). Other flatworm
species were taken from the flatworm collection established by the lab of Jochen Rink
(CBG, Dresden).
Large worms kept in isolation increasingly tend to fission, while social crowding is
known to reduce fissioning rates (48, 161). In order to prevent fissioning during long
term measurements but still track individual worms, we added small worm pieces in the
dishes of large worms. In long term experiments, we occasionally also added antibiotics
(ciprofloxacin, gentamicin) to ensure that the worms stay healthy during the entire
time course.
C.2. Image acquisition
Flatworms were placed one at a time into a plastic petri dish (90 mm) and filmed using
a Nikon macroscope (AZ 100M, 0.5x objective) and a Nikon camera set-up (DS-Fi1,
frame rate 3 Hz, exposure time 6 ms, total observation period 15 s, resolution 1280 x 960
pixel, conversion factor 44 pixel/mm). The movies were converted from AVI (provided
by the Nikon imaging software) to MP4 using Handbreak, reducing the bit-size by a
factor of 30 and facilitating further processing. For this purpose, we wrote a bash script
which calls HandbrakeCLI (the command line interface version) and transcodes the AVI
files using the H.264 standard for video compression and fixed optimized settings. We
carefully tested the conversion procedure to exclude distorting effects on the data. It
turns out that the worm tracking and shape analyses are even more robust if we first
transcode the files in comparison to loading the AVI files directly into the MATLAB
program.
129
C. WORM HANDLING AND MEASUREMENTS OF SIZE AND
SHAPE
For the size measurements, we aimed to obtain movies in which the worms are moving
rather straight and stretched out. This behavior corresponds to a flight response which
can be seen after the worms have been exposed to light. After the worms are placed
in the center of the petri dish, they try to escape the illumination of the macroscope
by moving towards the slightly darker region near the rim of the dish. The stage
is adjusted such that the worms traverse the field of view and are imaged just after
an initial acceleration phase. Student helpers have been thoroughly trained at the
macroscope to strictly adhere to the fixed imaging protocol.
C.3. Extracting size and shape from worm movies
Movies were analyzed off-line using custom-made MATLAB software, see Fig. C.1. A
first shape proxy was determined from movie frames via edge detection using a canny-
filter, followed by a dilation-erosion cycle and a filling of the shape. If possible, the
shape was extracted twice for each frame -- with and without background correction.
By comparing the two results, the program can reduce errors from particles touching
the worm boundary. In a subsequent refinement step, the worm perimeter was adjusted
by finding the steepest drop in intensity along directions transverse to the perimeter
proxy. This procedure together with interpolation steps allowed us to extract the worm
shape with sub-pixel accuracy,
After the perimeter was determined, we were able to compute further quantities such as
the worm area. The midline was obtained by searching for points with equal distances
to both sides of the worm boundary. Due to the large body plasticity, the measured
quantities such as worm area or length can easily vary by up to 30% during normal,
rather stretched gliding motion. It requires an extensive imaging training and a large
number of worms to extract reliable and reproducible data. Often, we analyzed several
movies per worm and experimental condition. In contrast to the biological variability,
the analysis software works very robustly and can cope with different noise levels and
manyfold intensity variations.
130
C.3 Extracting size and shape from worm movies
Figure C.1.: Extracting size and shape information from worm movies.
131
D. Shape reconstruction
and worm bending
D.1. On the reconstruction of a closed worm outline
In order to reconstruct a worm shape from a given radial distance profile, we have to
determine the constant increment ds along the worm perimeter such that the outline
(D.1)
closes on itself. This corresponds to (cid:88)
increasing with ds for radially convex shapes. (cid:80)
monotonously increasing (cid:80)
αi = 2π
i
with the angle αi between two consecutive radii ρi and ρi+1. The angle αi is monotonously
i αi can even be monotonously in-
creasing with ds if the shapes are not strictly radially convex but close to it. For a
i αi, there exists one choice for ds such that Eq. D.1 is
fulfilled. Small variations of the radial distance profile yield similar results for ds.
D.2. On the reconstruction of head shapes
The head perimeter does not completely close on its own.
Instead of Eq. D.1, the
following condition has to be fulfilled(cid:88)
i
αi = 2π − α0 ,
(D.2)
including the opening angle α0 towards the trunk. The angle α0 can be determined un-
der the assumption of an approximately symmetric head shape. For N radial distances
ρi describing the head, we obtain
α0 = arccos(ρN/2/ρ1) + arccos(ρN/2/ρN ) .
(D.3)
132
E. On growth and cell turnover
E.1. Signatures of aging in flatworms
Large and small worms are not only different in size but also show size-dependent
changes in system behavior, which have been previously related to characteristics of
more adult or juvenile animals, respectively (15, 18, 71, 111, 133). For this, we need
to distinguish between two different notions of "aging". The first one refers to the
progression of the physical time since birth. Secondly, there would be physiological
aging that is determined by varying characteristics of the body such as sexual maturity
and also symptoms of decline as well as an increased mortality. Aging in the second
sense might potentially be independent of the life time of the organism and might
instead rather depend on its size or environmental conditions.
For sexual flatworms, it has in fact been reported that growth, metabolic rate, regene-
ration abilities and reproduction change with age (1, 86, 92, 111). Additionally, aging
is often attributed to a decreasing length of telomeres (i.e. the end regions of the
chromosomes) with each cell division. Indeed, this telomere erosion has been described
in sexual Smed and it correlates with the total life time of the worm (213). In contrast,
other features seem to rather depend on size instead of the life time. For example,
sexual flatworms develop reproductive organs when growing but also reabsorb their
germ line cells again when shrinking during starvation periods (30, 33, 140, 193).
For asexual worms, as we consider in the experiments, aging effects are even more
difficult to discuss. There is no well-defined birth and, at most, fissioning could be
considered as death and birth at the same time. Nevertheless, asexual Smed still show
size-dependent physiological changes. For example, the fissioning rate increases for
large worms (215) and precursor cells of the germ line, which still exist in asexual
flatworms, show a similar size-dependence as the germ line in sexual worms (84, 230).
Telomere erosion has also been observed in asexual Smed , yet the length is restored
during regeneration in contrast to the sexual strains (213). In this sense, physiological
aging in asexual worms might be considered to be reversible and regeneration might be
interpreted as a rejuvenation event, which also includes the renewal of many somatic
133
E. ON GROWTH AND CELL TURNOVER
cells during remodeling (50, 86). It is an interesting question whether the worms also
show a physiologically rejuvenation while starving (86, 91, 140).
In summary, there is some evidence that sexual worms show irreversible signatures of
aging with progression of their life time while asexual strains might be able to return
to a more juvenile state. As a sidenote, sexual worms survive longer than asexual
worms after irradiation (229), which could be due to a reduced turnover rate. One
might speculate that sexual worms rather invest in producing a germ line on the cost
of body maintenance. Such a potentially insufficient replacement of damaged cells
and the corresponding deterioration of function alongside with a higher mortality can
be considered as a hallmark of aging (158). However, most importantly, the discussion
shows how difficult it is to unambiguously characterize aging, especially as it is in general
lacking a clear definition (53). It will be left for future works to investigate whether
size-dependent changes in flatworms might be linked to aging and rejuvenation.
E.2. Additional size measurements and growth dynamics
In this section, we provide more details and additional measurements of worm sizes and
growth dynamics to back up our discussion in Section 5.2.
We define the worm area for a particular measurement as the average over the 10 frames
with the largest values. Scaling laws are fitted by a robust algorithm with bi-squared
weights. Thus, the spread of the data is assumed to correspond to the measurement
uncertainty. This circumvents the difficulties with estimating the error for each data
point from the size measurements. Fig. E.1(a) shows the scaling relation between area
and length for well-fed worms (3 days after feeding). The fit yields a scaling exponent
of 1.69 ± 0.01, which is slightly lower than for the starving worms in Fig. 5.3(a).
As a byproduct of the cell number measurements using histones, Albert Thommen also
determined the protein mass in the worms. Even though the measurements are not
sufficiently calibrated to obtain absolute numbers, the respective scaling laws might
still be meaningful. While the total mass per cell seems to change with worm size, the
protein content per cell appears to stay rather constant, see Fig. E.1(b). It remains to
validate this result with additional experiments and to investigate whether it could be
linked to potential lipid stores.
Oviedo et al. have counted the number of cintillo cells around the head margin (150). If
we assume an equidistant positioning of these sensory cells and approximate the head
by a semi-circle, the width of the worm head can be considered to be proportional to
the number of cintillo cells, see Fig. E.1(c). The resulting scaling law for the width
134
E.2 Additional size measurements and growth dynamics
Figure E.1.: (a) Relationship between area and length in well-fed worms. The red
curve represents the result of a robust fit in the double-logarithmic plot (inset). In
comparison, the dashed line denotes the scaling law of starved worms with exponent
1.8 (Imaging by Nicole Alt under the supervision of the author, analysis by the author
includes 281 measurements 3 days after feeding). (b) Protein content per cell does not
depend on worm size (measurement by Albert Thommen, analysis by the author, 45
worms, absolute values are only approximate). (c) Scaling of the number of cintillo
cells along the head margin with worm length. Data measured by Oviedo et al. (150).
(d) Feeding response after two weeks of starvation (Imaging by Ian Smith under the
supervision of the author, analysis by the author, 20 worms).
135
051015202501020304050Area in mm2Length in mmLog lengthLog area(a)(b)0123456700.20.40.60.81Cell number in 106Protein content per cell in ng0246810020406080Number of cintillo cellsLength in mm(c)(d)0135790.70.80.911.11.2Time after feeding (days)Normalized areaE. ON GROWTH AND CELL TURNOVER
Figure E.2.: (a) Growth dynamics in Girardia tigrina based on measurements by
Bagun`a et al. (15, 18). (b) The plot of our data for Smed corresponding to the plot
in (a) shows a similar trend for the growth rates (blue) but a different behavior for
the degrowth rates (red). (c) Growth and degrowth rates extracted from the data of
Oviedo et al. for Smed are shown as black curves (150). These curves (for starvation
and feeding twice a week) show a similar trend to our data (for starvation and feeding
once a week), yet they do not agree quantitatively.
with an exponent of 0.67 ± 0.03 is in agreement with our measurements of area versus
length.
Upon feeding, worms show a characteristic growth response. The immediate increase
in worm size is mainly due to stuffing as discussed in Section 5.2.2. When comparing
the feeding response after two weeks of starvation with the peak after one week of
starvation of Fig. 5.4, we see no obvious dependence on feeding history.
When being fed regularly, worms grow and degrow depending on the feeding frequency.
Bagun`a et al. have determined the rates of addition and removal of cells in Girardia
tigrina for various feeding conditions (15, 18). From this data, we can back-calculate
the growth rates which they originally measured, see Fig. E.2(a). While the growth
trends are similar to our results, the degrowth rates seem not to vary with size in
contrast to our data, see Fig. E.2(b).
Oviedo et al. have obtained linearly increasing and decreasing functions of the worm
length with time during feeding and starvation, respectively (150). We can compute
growth and degrowth rates from their data sets and compare them to our measurements
in terms of worm length, see Fig. E.2(c). The linear growth behavior yields L/L ∝
±1/L. Thus, the absolute rates are decreasing for large worms in agreement with our
results. However, the curves do not fit our data points very well. They are shifted
downwards even though the growing worms are even fed twice a week. In particular,
the absolute values for the growth rate are smaller than for the degrowth rate, which
contradicts our observation.
136
Length in mm24681012Growth rate (cell number) in %/ day-10122x/week1x/week2x/month1x/monthstarvation05101520−4−20246Length in mmGrowth rate (cell number) in %/day051015−2−101234Length in mmGrowth rate (length) in %/day(b)(c)(a)1x/weekstarvationstarvation2x/weekE.3 Measuring cell cycle times
Figure E.3.: Dividing stem cells progress through four distinct cell cycle stages, which
can be identified by respective markers. DNA replication happens in S-phase and cells
divide in M-phase, while G1 and G2 are merely resting phases including check points
for error control. We discuss an experiment to measure the cell division rates by
blocking the transition from M-phase to G1 (red cross).
E.3. Measuring cell cycle times
Stem cells move through four distinct cell cycle phases, which can be characterized by
different biochemical markers: pcna (all stem cells), clumping pcna (S-phase), cycling
B (G2-phase), H3P (M-phase), see Fig. E.3. An increase of H3P marker has been
measured upon feeding (71). Under the assumption that the duration of the M-phase is
relatively constant, this can be interpreted as an increase in cell division rate. However,
one cannot extract actual rates from this measurement if one does not know the duration
of the M-phase.
Bagun`a et al. have performed experiments from which they obtained cell division
rates (12, 13, 15, 18, 176, 178). They blocked the progression through M-phase using
colchicine as indicated by the red cross in Fig E.3, which results in an increasing number
of cells in M-phase. Under the assumption that the M-phase has a constant duration
and cell death can be neglected, the time derivative of the number of accumulating cells
in M is a direct read-out of cell division rate. We plan to perform similar experiments
using the cell cycle markers discussed above, which have not been available for the
previous works. In particular, we aim to reveal how the division rates depend on size
and feeding. By subtracting division rates from growth rates, we will also be able to
extract the corresponding cell loss rates. Thus, we hope to reveal whether the control
points of cell turnover are associated with the mechanisms for the addition or loss of
cells.
Furthermore, the time scale of the first response upon feeding might also hint at how the
cell division rate is controlled. It has been discussed that there might be a population
of slowly cycling cells that mainly stay in G2 to allow a fast regeneration and feeding
137
SMG1G2 celldifferentiation(H3P)(clumped pcna)(cyclin B)E. ON GROWTH AND CELL TURNOVER
response (12, 13, 139, 142, 170, 182). This represents a major investment of cellular
resources as G2 cells already have a duplicated DNA. It would also be in contrast to
the arrest in G1, typically observed for other organisms. If we were to measure a very
fast division response, it would support the hypothesis of G2 arrested cells.
E.4. Measuring cell turnover dynamics on the organism level
Here, we discuss the theoretical framework for measuring turnover dynamics on the
organismal level. It will enable us to test and compare various models with specific
assumptions on how cell division and loss rates depend on feeding and worm size.
The experimental approach relies on Histone labeling by heavy isotopes (SILAC). It
is inspired by the measurements of neurogenesis dynamics in adult humans exploiting
the known C14 concentration in the atmosphere due to nuclear bomb tests (203). First,
we present the main principles of this paper. Next, we apply the idea to flatworms
suggesting an adapted experimental protocol. Finally, we discuss details of histone
labeling.
E.4.1. Measurement of C14 reveals dynamics of neurogenesis in humans
First, we briefly review the main ideas of the work by Spalding et al. (203) and introduce
our notation. The paper describes a measurement of the turnover of hippocampal
neurons in humans. It is based on the known C14 concentration in the atmosphere due
to nuclear bomb tests, which becomes incorporated into the DNA as a label when a new
cells is made. From this, we can compute the total C14 concentration in the DNA of a
person of a particular age if we assume a particular model for the turnover dynamics.
This result can then be benchmarked using the measured C14 in dead people.
The starting point is that we have a list of several possible scenarios for how cell birth
and cell death depends on each other as well as on the age of the cells and the age of
the person. For each of these scenarios, we can compute a distribution n(A, a) that
describes the number of cells of age a in a person of age A. This distribution changes
over time because of aging (left) and because of cell death (right)
∂A n(A, a) + ∂a n(A, a) = −kloss(A, a) n(A, a) ,
(E.1)
where kloss is the loss rate which might depend on the age of the person as well as the
age of the cell. Cell birth is considered by the boundary condition
n(A, 0) = n0(A) ,
(E.2)
138
E.4 Measuring cell turnover dynamics on the organism level
Figure E.4.: Schematic representation of the C14 experiment for the measurement of
turnover dynamics in human neurons (203).
where n0(A) is the influx of cells due to cell division. Cell birth might depend on the
age of the person explicitly but also implicitly by e.g. depending on the number of dying
cells at each time point.
Furthermore, as an initial condition it is assumed that a first set of cells is made in the
embryo half a year before the person is born, resulting in an initial condition
n(−0.5, a) = δ(a) .
(E.3)
This includes a delta function δ (which is zero for a (cid:54)= 0), describing the fact that cells
are born with age 0.
By solving the system of Eqs. E.1 - E.3, we determine the number of cells n(A, a) of
age a for any age A of the person, given a specific model for kloss(A, a) and n0(A). In
particular, we can predict the age distribution n(Ad, a) in a person who died at age Ad.
For a person who died in year td, we can compute the total concentration of C14 label
by
(cid:90) Ad+0.5
C(td, Ad) =
0
N (Ad)
Cin(td − a) n(Ad, a)
da ,
(E.4)
if the concentration Cin(t) of C14 inside the body (which could thus be incorporated
in new cells) is known at any time point. The integral sums over all cells of all ages,
multiplied by the amount of label Cin at the respective time of birth td − a.
The result is normalized by the total number of all considered cells in a dead person,
which is given by N (Ad) =(cid:82) Ad+0.5
0
n(Ad, a) da. The upper bound of integration reflects
the maximum age obtained by cells, which persist during the full life span of the person.
Note that the summand of 0.5 was missing in the original publication.
139
Knowing time course of C14 concentrationin the bodyGuessing models for the rates of cell death & birthCalculating total C14 label in a person that died at age in year :Comparing tomeasurementof Rating modelsEq.G.1-G.4E. ON GROWTH AND CELL TURNOVER
The concentration Cin of C14 inside the body is estimated from the known C14 concen-
tration in the atmosphere by introducing a heuristic time lag of one year. Variations
in the time lag are claimed not to significantly change the result.
Different turnover models can be rated using the Akaike Information Criterion (AIC),
which avoids over-fitting when comparing models with a different number of parameters.
In brief, this technique calculates the likelihood that a certain model has generated the
measured data. As a more complex model with more parameters tends to fit the data
better, it introduces a simple but mathematically well-justified penalty term for the
number of parameters based on information theory.
E.4.2. Adapting the C14 technique to cell turnover in flatworms
In contrast to the neurogenesis paper, we need to consider a different label than the C14
concentration in the atmosphere for flatworms. We choose to use the SILAC protocol
(stable isotope labeling of amino acids) (32). The worms are fed with mouse liver,
in which more than 96% of the amino acid lysine is replaced by the heavy isotope
form 13C6-lysine.
and in particular into histones, which package the DNA in the nucleus. Histones are
In consequence, the labeled lysine gets incorporated into proteins
believed to be only synthesized when a cell is made and to remain stable throughout
the lifetime of the cell. Thus, the fraction of labeled histones can serve as a read-out
for cell turnover.
The technique from the neurogenesis paper has to be modified due to the unknown initial
age distribution. -- In flatworms, we encounter two main differences: On the positive
side, we can directly measure (and to some extent even deliberately vary) the label
content in the body Cin(t). On the negative side, we do not know an initial age
distribution analogous to Eq. E.3. In the following, we suggest experimental protocols
that resolve this problem.
Similar to Eq. E.1-E.2, we can describe the time evolution by
∂t n(t, a) + ∂a n(t, a) = −kloss(t, a) n(t, a)
n(t, 0) = n0(t) .
(E.5)
(E.6)
This set of equations can be solved analytically
n(t, a) = n0(t − a) exp
−
.
(E.7)
(cid:90) a
kloss(t − a + a(cid:48), a(cid:48)) da(cid:48)(cid:19)
(cid:18)
0
140
E.4 Measuring cell turnover dynamics on the organism level
The number of cells of age a at time t is the same number that was born at time t − a,
but reduced by all the death events in between, described by the exponential term with
the integral over the cell death time course. If we knew or were able to model all the
cell birth and death processes for all times in the past, this solution would give us the
full age distribution at time t. Alternatively, one could only consider birth and death
rates starting from a particular time point, if one knew the initial distribution at this
time, see Eq. E.3.
Unfortunately, in flatworms we do neither know the age distribution at one initial time
point nor the full feeding and growth history, which would enable us to model n0 and
kloss. Luckily, there are two options to solve this problem, each relies on a particular
assumption.
First option: Assuming a stem cell control model. -- If there is no feedback from
kloss(t, a) or n(t, a) on cell birth n0(t) (i.e. mainly negligible apoptosis-induced cell
division), we can determine the age distribution n(t, a) at least partially for all ages
a < t − t0, assuming the experiment (i.e. the well-controlled feeding conditions) starts
at t0. The longer we perform the experiment, the more of the distribution we can
compute. As cells do not live forever and the number of old cells decays exponentially
(see Eq. E.7), we will approximately obtain the full age distribution with time.
Second option: Assuming the system relaxes to a steady state for a reference feeding
condition. -- If we provide a constant feeding condition, for which the worms neither
grow nor degrow, we might assume that the age distribution eventually relaxes to a
steady state n∗(a), which is a solution to the equation
This equation can be solved for a particular choice of kloss(a) and n∗
0:
0 e−(cid:82) a
n∗(a) = n∗
∂a n∗(a) = −kloss(a) n∗(a) .
0 kloss(a(cid:48)) da(cid:48)
.
(E.8)
(E.9)
In order to obtain such a steady state, we would need to constantly provide the worms
with a small and well-controlled amount of food. Even though, theoretically it would
be the most clean approach, experimentally it is rather difficult to frequently pipette a
well-defined amount of food to each worm.
Modified second option: Assuming a quasi-steady state for a reference feeding condition.
-- If worms are fed every second week, on average, they will also neither grow nor
degrow but show size oscillations with a frequency of two weeks.
Importantly, the
141
E. ON GROWTH AND CELL TURNOVER
steady state argument can be modified for such a periodically oscillating quasi-steady
state n∗(t, a). This quasi-steady state can be used as an initial condition for other
feeding schemes in analogy to Eq. E.3.
E.4.3. Label dynamics after a single feeding event
Here, we illustrate the measurement scheme by discussing the effect of a single feeding
pulse with labeled lysine in a simplified example. The total influx of the amino acid
lysine is given by
JAF =
e−t/τF ,
NAF
τF
(E.10)
where NAF is the total amount of the amino acid lysine in the gut after feeding and
τF is the digestion time of the food. A certain fraction ΦAF of the lysine in the food is
assumed to be labeled by heavy isotopes.
After feeding with labeled food, the fraction ΦA of the labeled amino acid lysine in-
creases in the animal. Lysine belongs to the building blocks of histones and other
proteins, which also will become labeled by the incorporation the amino acid. As a
simple example, let us assume there are only two proteins in the body that take up
lysine: a fast turnover protein and histone. The result is qualitatively the same, inde-
pendent of the number of proteins that incorporate lysine and might be turned over
at very different rates. We define the label fraction ΦH of lysine in histones as the
ratio between all labeled lysine in histones and the total amount of lysine in histones.
Analogously, we also define the label fraction of the lysine in the fast turnover protein
ΦP .
If we furthermore assume that cell death happens stochastically and does not depend
on the age of the cell, the dynamics of ΦA are captured by
(1 − ρH )NAH
NA
(1 − ρP )NAP
NA
(ΦAF−ΦA)+
JAF
NA
∂tΦA =
(ΦP−ΦA)JP +
(ΦH−ΦA) kloss N .
(E.11)
Here, NA denotes the total amount of lysine in the worm, (1 − ρP ) is the fraction
of lysine that can be recycled from fast turnover proteins and NAP is the amount of
lysine per protein. Analogously, (1 − ρH ) is the fraction of lysine that can be recycled
from histones and NAH is the amount of lysine in all the histones in a cell. The label
fraction of the amino acid only changes via recycling or feeding. Each further protein
that incorporates lysine would be represented by an analogous term.
142
E.4 Measuring cell turnover dynamics on the organism level
Figure E.5.: (a) Example dynamics for a single labeling pulse. The label fraction of a
fast turnover protein (dashed green) acts as an approximate read-out of the available
amount of labeled amino acid lysine (gray), which can be incorporated into other
proteins such as histones (red). As an example, we have chosen the fast turnover half as
fast as the label digestion and ten times as fast as the cell turnover. If digestion and the
fast turnover happen at comparable time scales, the gray and the green curves collapse.
(b) Labeling scheme: the label dynamics in the fast turnover protein corresponds to
labeled lysine in the worm, while the label incorporation in histones reflects the cell
turnover.
Similarly, the dynamics of the label fractions of the fast turnover proteins and the
histones are
∂tΦP =
∂tΦH =
(ΦA − ΦP ) JP
(ΦA − ΦH ) n0 .
NAP
NP
NAH
NH
(E.12)
(E.13)
The label fractions of lysine ΦA, the fast turnover protein ΦP and histone ΦH are
shown in Fig. E.5(a) for a single feeding pulse with labeled food. ΦA (gray) cannot be
measured directly, yet the fast turnover protein (dashed green) immediately follows and
can be used as a read-out of labeled lysine. Finally, the histones (red) also incorporate
the label on a much longer time scale given by the life times of the cells. Fig. E.5(b)
sketches the main principle of the experiment.
By combining the measured fraction of labeled fast turnover proteins with a model on
cell turnover, we can compute
ΦH =
(cid:90) ∞
ΦP (t − a) n(t, a)
0
N (t)
da
(E.14)
and compare it to the measured value of ΦH (t). Here, ΦP (t) takes over the role of
Cin(t) in Eq. E.4. Yet, in contrast to the neurogenesis paper, the concentration of label
143
TimeLabel fractionAmino acidFast turnover proteinHistone(a)(b)feedingamino acid lysinefast turnover proteinhistoneclose to equilibriumtime scale of cell turnover(read out of labeled amino acid)E. ON GROWTH AND CELL TURNOVER
in the worm (which can be incorporated into histones) can be directly measured by
monitoring the label in high turnover proteins.
E.4.4. Challenges and advantages of the application to flatworms
Lacking knowledge about an initial age distribution, we have to rely on experimental
protocols and specific assumptions that eliminate a potential history-dependence. We
can tackle this problem by assuming a periodically varying quasi-steady state when
feeding every second week and maybe independently measure the respective cell birth
rate. We will be able to estimate the time scale of the relaxation to this quasi-steady
state based on the spreading of neoblast clones in irradiated worms. Once it has been
characterized, the quasi-steady state can act as the initial configuration for other feeding
schemes.
As an advantage in flatworms, we might be able to deliberately vary the time course of
the internalized label by alternating between labeled and non-labeled food. This might
help to distinguish different models for birth rate n0 and death rate kloss to a higher
resolution. Note that the more Cin(t) or ΦP (t) varies with time, the more details of
the models can be resolved. If the internal label was kept constant, almost any model
would fit the data equally well. As a second advantage, we can directly determine the
label in the body using high turnover proteins.
144
List of Figures
1.1. Development and Regeneration . . . . . . . . . . . . . . . . . . . . . . .
1.2. Body axes and morphogenesis . . . . . . . . . . . . . . . . . . . . . . . .
1.3. Morphogen gradients . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4. Gene expression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.5. Complexity and regneration capabilities . . . . . . . . . . . . . . . . . .
1.6. Organs of Smed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.7. Growth and regeneration of Smed
. . . . . . . . . . . . . . . . . . . . .
2
4
6
7
9
10
12
1.8. Sequential development vs. regeneration from arbitrary initial conditions 14
1.9. French flag model and Turing mechanism . . . . . . . . . . . . . . . . .
1.10. Two possible topologies with two Turing molecules . . . . . . . . . . . .
2.1. Perfect versus approximate shape scaling of morphogen gradients . . . .
2.2. Gradient scaling and implications for the amplitude
. . . . . . . . . . .
2.3. Scaling with an autonomously controlled expander . . . . . . . . . . . .
2.4. Expander-repression model and deviations from scaling
. . . . . . . . .
2.5. Two main classes of expander models . . . . . . . . . . . . . . . . . . . .
2.6. Hierarchy of Turing patterns
. . . . . . . . . . . . . . . . . . . . . . . .
2.7. Relative source size of a Turing system . . . . . . . . . . . . . . . . . . .
2.8. Absence of scaling in classcial Turing patterns . . . . . . . . . . . . . . .
3.1. Turing patterns vs. scaling patterns . . . . . . . . . . . . . . . . . . . . .
3.2. Scaling in a Turing system with expander feedback . . . . . . . . . . . .
3.3. Fixed points and dynamics in a Turing system with expander feedback .
3.4. Amputation experiment with ten cuts
. . . . . . . . . . . . . . . . . . .
3.5. β-catanin gradient in flatworms . . . . . . . . . . . . . . . . . . . . . . .
4.1. Flatworm speed and phenotypes
. . . . . . . . . . . . . . . . . . . . . .
4.2. Three shape modes characterize body shape dynamics . . . . . . . . . .
4.3. Distinguishing head morphologies of flatworm species.
. . . . . . . . . .
15
18
25
26
29
31
33
36
37
39
43
45
47
51
55
61
64
67
5.1. Cell turnover and growth . . . . . . . . . . . . . . . . . . . . . . . . . .
71
145
LIST OF FIGURES
. . . . . . . . . . . . . . . . . . .
5.2. Stem cell control vs. cell death control
. . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3. Allometric scaling Laws
5.4. Growth response upon feeding . . . . . . . . . . . . . . . . . . . . . . . .
5.5. Growth and degrowth rates . . . . . . . . . . . . . . . . . . . . . . . . .
5.6. Non-trivial growth dynamics in flatworms . . . . . . . . . . . . . . . . .
5.7. Dynamic energy storage . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.8. Growth models with energy storage . . . . . . . . . . . . . . . . . . . . .
5.9. Catabolic default state and mitotic bursts upon feeding . . . . . . . . .
5.10. Turnover of epidermis cells
. . . . . . . . . . . . . . . . . . . . . . . . .
5.11. Measuring cell turnover dynamics with two labels . . . . . . . . . . . . .
72
74
78
80
82
86
88
92
94
97
A.1. Eigenvalues of the reaction-diffusion matrix . . . . . . . . . . . . . . . . 108
A.2. Homogeneous steady state of our Turing system . . . . . . . . . . . . . . 113
A.3. (1, 1)-pattern and its source size . . . . . . . . . . . . . . . . . . . . . . . 115
A.4. Constructing a hierarchy of Turing patterns . . . . . . . . . . . . . . . . 118
A.5. Linear stability analysis of inhomogeneous Turing patterns . . . . . . . . 119
A.6. RNAi experiments in Turing systems with expander
. . . . . . . . . . . 124
C.1. Extracting size and shape from worm movies
. . . . . . . . . . . . . . . 131
E.1. Additional scaling laws and feeding peaks . . . . . . . . . . . . . . . . . 135
E.2. Previously measured growth dynamics in flatworms . . . . . . . . . . . . 136
E.3. Cell division rates from cell cycle dynamics
. . . . . . . . . . . . . . . . 137
E.4. Measurement of turnover dynamics in humans . . . . . . . . . . . . . . . 139
E.5. Measuring turnover dynamics on the organism scale in flatworms . . . . 143
146
List of Variables
In Chapter 2. Scaling in morphogen systems,
and Chapter 3. Scaling and regeneration of self-organized patterns:
C
C0
Z
ν
w
α
β
D
λ =(cid:112)D/β
E
νE
wE
αE
βE
DE
λE
Cth
β0
Eth
χλ = λ/L
χw = w/L
A, B
Bc
a, b
RA, RB
P
h
(cid:96) = (cid:104)P(cid:105) L
αA, αB
βA, βB
DA, DB
λA, λB
M
s
q
m
concentration profile of morphogens
amplitude of morphogen profile
shape function of morphogen profile
production function of morphogen
source size of morphogen
production rate of morphogen
effective degradation rate of morphogen
effective diffusion coefficient of morphogen
length scale of morphogen profile
concentration profile of expander
source function of expander
source size of expander
production rate of expander
effective degradation rate of expander
effective diffusion coefficient of expander
length scale of expander profile
threshold for expander production (expansion-repression model)
morphogen degradation in the absence of expander (exp.-rep. model)
conc. threshold for feedback of expander on morphogen degradation
proportionality factor between λ and L
proportionality factor between w and L
concentration profiles of Turing molecules (activator and inhibitor)
formal shift of inhibitor concentration between the two Turing topologies
small perturbations of steady state profiles
generic reaction functions of Turing system
switch-like source function of Turing system (Hill or Heaviside function)
exponent for Hill function
size of self-organized source region (includes spatial average of P )
production rates of Turing molecules
degradation rates of Turing molecules
diffusion coefficients of Turing molecules
characteristic length scales of Turing molecules
linearized reaction-diffusion matrix
wavenumber of Fourier modes
eigenvalues of reaction-diffusion matrix
number of contiguous source regions
147
LIST OF FIGURES
number of source regions touching the system boundary
n
σ = 2m − n pattern number
Lσ
κA, κB, κE
E0
f(m,n)
g
χ = B/A
system size beyond which a Turing pattern of pattern number σ exists
coupling constant for degradation rates in scalable Turing system
hypothetically imposed expander level in scalable Turing system
source size on the nullcline of scalable Turing system for ∂tE = 0
source size of a homogeneous dynamic state for low expander levels
constant ratio in a homogeneous dynamic state for low expander levels
In Chapter 4. Flatworm shape dynamics and motility:
arc-length along worm circumference
total length of circumference
normalized arc-length
radial position vector
center point of worm
radial distance
mean radius
normalized radius profile defining the shape
average shape variations
number of worm outlines
number of data points along worm perimeter
s
L
s = s/L
r
r0
ρ = r − r0
ρ
ρ = ρ(s)/ρ
ρ0
Nw
Nr
Ri,j = ρi(sj) data matrix
C
v
B
covariance matrix
eigenvectors or shape modes
eigenvalues or shape scores
In Chapter 5. Quantitative study of flatworm growth and cell turnover:
N
kdiv
kloss
K = kdiv − kloss
Et
ec = E/N
Jf
jf = jf /N
µ
K0
es
τK
e0
a
n
N(cid:96)
total cell number
average cell division rate
average cell loss rate
average growth rate
total stored metabolic energy in the worm
stored metabolic energy per cell
influx of metabolic energy due to feeding
influx of metabolic energy per cell
metabolic consumption rate per cell
degrowth rate in the absence of any stored energy (model 1)
energy threshold for the switch between growth and degrowth (model 1)
relaxation time of the growth rate to a quasi steady state (model 2,3)
target value of the metabolic energy (model 2,3)
age of cells
number density of cells of age a
number of labeled cells
148
LIST OF FIGURES
ki, kd, kr
¯kd
¯kd(cid:96), ¯kr(cid:96)
φ(cid:96)
N1, N2
¯kd1, ¯kr1
¯kd2, ¯kr2
φ12 = N1/N2
age-dependent rates of insertion, deletion and replacement of cells
averaged rate of deletion across the tissue
averaged rates of deletion and replacement of labeled cells
fraction of labeled cells
number of cells with one or two labels
averaged rates of deletion and replacement of cells with one label
averaged rates of deletion and replacement of cells with two labels
label ratio
149
References
[1] Abeloos, M. (1929). Recherches exp´erimentales sur la croissance et la r´eg´en´eration chez
les planaires. Ph.D. thesis, Paris. 79, 133
[2] Abrams, E.W. & Mullins, M.C. (2010). Early zebrafish development:
maternal genes. Current opinion in genetics & development, 19, 396 -- 403. 13
It's in the
[3] Adell, T., Sal`o, E., Boutros, M. & Bartscherer, K. (2009). Smed-Evi/Wntless
is required for beta-catenin-dependent and -independent processes during planarian re-
generation. Development (Cambridge, England), 136, 905 -- 910. 12, 54, 99
[4] Adell, T., Cebri`a, F. & Sal´o, E. (2010). Gradients in planarian regeneration and
homeostasis. Cold Spring Harbor perspectives in biology, 2, a000505. 6, 9, 12, 13, 54, 57,
98
[5] Affolter, M. & Basler, K. (2007). The Decapentaplegic morphogen gradient: from
pattern formation to growth regulation. Nature reviews. Genetics, 8, 663 -- 74. 5
[6] Agrawal, N., Dasaradhi, P.V.N., Mohmmed, A., Malhotra, P., Bhatnagar,
R.K. & Mukherjee, S.K. (2003). RNA Interference: Biology, Mechanism, and Appli-
cations. Microbiology and Molecular Biology Reviews, 67, 657 -- 685. 8
[7] Agutter, P.S. & Wheatley, D.N. (2004). Metabolic scaling: consensus or contro-
versy? Theoretical biology & medical modelling, 1, 13. 87, 89
[8] Almuedo-Castillo, M., Sal´o, E. & Adell, T. (2011). Dishevelled is essential for
neural connectivity and planar cell polarity in planarians. Proceedings of the National
Academy of Sciences, 108, 2813 -- 2818. 9, 58, 59, 62, 66
[9] Almuedo-Castillo, M., Sureda-G´omez, M. & Adell, T. (2012). Wnt signaling
in planarians: new answers to old questions. The international journal of developmental
biology, 56, 53 -- 65. 6, 12, 13, 54, 57, 62, 98
[10] Almuedo-Castillo, M., Crespo, X., Seebeck, F., Bartscherer, K., Sal`o, E. &
Adell, T. (2014). JNK Controls the Onset of Mitosis in Planarian Stem Cells and Trig-
gers Apoptotic Cell Death Required for Regeneration and Remodeling. PLoS Genetics,
10. 12
[11] Averbukh, I., Ben-Zvi, D., Mishra, S. & Barkai, N. (2014). Scaling morphogen
gradients during tissue growth by a cell division rule. Development, 141, 2150 -- 6. 24, 27,
42, 98, 102
150
REFERENCES
[12] Bagun`a, J. (1974). Dramatic mitotic response in planarians after feeding, and a hypoth-
esis for the control mechanism. Journal of Experimental Zoology, 190, 117 -- 122. 77, 91,
137, 138
[13] Bagun`a, J. (1976). Mitosis in the Intact and Regenerating Planarian Dugesia mediter-
ranea n.sp. I. Mitotic studies during growth, feeding and starvation. Journal of Experi-
mental Zoology, 195, 53 -- 64. 12, 75, 76, 77, 83, 91, 137, 138
[14] Bagun`a, J. (1976). Mitosis in the Intact and Regenerating Planarian Dugesia mediter-
ranea n.sp. II. Mitotic studies during regeneration, and a possible mechanism of blastema
formation. Journal of Experimental Zoology, 195, 65 -- 79. 12
[15] Bagun`a, J. (2012). The planarian neoblast: the rambling history of its origin and some
current black boxes. The International journal of developmental biology, 19 -- 37. 3, 10, 11,
79, 133, 136, 137
[16] Bagun`a, J. & Romero, R. (1981). Quantitative analysis of cell types during growth,
degrowth and regeneration in the planarians Dugesia mediterranea and Dugesia tigrina.
Hydrobiologia, 84, 181 -- 194. 10, 71, 75, 76, 77, 83, 91
[17] Bagun`a, J., Sal´o, E. & Auladell, C. (1989). Regeneration and pattern formation
in planarians. III. Evidence that neoblasts are totipotent stem cells and the source of
blastema cells. Development, 107, 77 -- 86. 10
[18] Bagun`a, J., Romero, R., Sal´o, E., Collet, J., Auladell, C., Ribas, M., Ri-
utort, M., Garc´ıa-Fern`andez, J., Burgaya, F. & Bueno, D. (1990). Growth,
degrowth and regeneration as developmental phenomena in adult freshwater planarians.
In H.J. Marthy, ed., Experimental embryology in aquatic plants and animals, 129 -- 162,
Plenum Press, New York. 10, 79, 83, 133, 136, 137
[19] Barkai, N. & Ben-Zvi, D. (2009). 'Big frog, small frog' -- maintaining proportions in
embryonic development. The FEBS journal , 276, 1196 -- 1207. 29
[20] Barkai, N. & Shilo, B.Z. (2009). Robust generation and decoding of morphogen gra-
dients. Cold Spring Harbor perspectives in biology, 1, 1 -- 15. 24
[21] Barnard, C. (2004). Animal Behaviour. Mechanism, Development, Function, and Evo-
lution. Pearson Education Canada. 59
[22] Beane, W.S., Morokuma, J., Adams, D.S. & Levin, M. (2012). A Chemical Genetics
Approach Reveals H,K-ATPase-Mediated Membrane Voltage is Required for Planarian
Head Regeneration. Chemical Biology, 18, 77 -- 89. 57
[23] Beane, W.S., Morokuma, J., Lemire, J.M. & Levin, M. (2013). Bioelectric sig-
naling regulates head and organ size during planarian regeneration. Development (Cam-
bridge, England), 140, 313 -- 22. 13, 57
[24] Ben-Zvi, D. & Barkai, N. (2010). Scaling of morphogen gradients by an expansion-
repression integral feedback control. Proceedings of the National Academy of Sciences of
151
REFERENCES
the United States of America, 107, 6924 -- 9. 22, 23, 24, 26, 27, 29, 30, 31, 32, 33, 41, 42,
56
[25] Ben-Zvi, D., Shilo, B.Z., Fainsod, A. & Barkai, N. (2008). Scaling of the BMP
activation gradient in Xenopus embryos. Nature, 453, 1205 -- 11. 27, 57
[26] Ben-Zvi, D., Pyrowolakis, G., Barkai, N. & Shilo, B.Z. (2011). Expansion-
repression mechanism for scaling the Dpp activation gradient in Drosophila wing imaginal
discs. Current biology, 21, 1391 -- 6. 24, 26, 27, 29, 30, 31, 33, 41, 56, 98
[27] Ben-Zvi, D., Shilo, B.Z. & Barkai, N. (2011). Scaling of morphogen gradients. Cur-
rent opinion in genetics & development, 21, 704 -- 10. 22, 24, 26, 27, 42
[28] Benazzi, M., Ballester, R., Bagun´a, J. & Puccinelli, I. (1972). The Fissiparous
Race of the Planarian Dugesia Lugubris S.L. Found in Barcelona (Spain) Belongs to the
Biotype G: Comparative Analysis of the Karyotypes. Caryologia, 25, 59 -- 68. 129
[29] Bergmann, S., Sandler, O., Sberro, H., Shnider, S., Schejter, E., Shilo, B.Z.
& Barkai, N. (2007). Pre-Steady-State Decoding of the Bicoid Morphogen Gradient.
PLoS Biology, 5, e46. 24
[30] Berninger, J. (1911). Uber die Einwirkung des Hungers auf Planarien. In J. Spengel,
ed., Zoologische Jahrbucher , 181 -- 216, Verlag Gustav Fischer, Jena, 30th edn. 133
[31] Bittig, T. (2008). Morphogenetic signaling in growing tissues. Ph.D. thesis, TU Dresden.
102
[32] Boser, A., Drexler, H.C.A., Reuter, H., Schmitz, H., Wu, G., Scholer, H.R.,
Gentile, L. & Bartscherer, K. (2013). SILAC proteomics of planarians identifies
Ncoa5 as a conserved component of pluripotent stem cells. Cell reports, 5, 1142 -- 55. 140
[33] Bowen, I.D., Ryder, T.A. & Dark, C. (1976). The effects of starvation on the pla-
narian worm Polycelis tenuis iijima. Cell and Tissue Research, 169, 193 -- 209. 133
[34] Calow, P. (1977). The joint effect of temperature and starvation on the metabolism of
triclads. Oikos, 29, 87 -- 92. 81
[35] Cebri`a, F. & Newmark, P.A. (2005). Planarian homologs of netrin and netrin receptor
are required for proper regeneration of the central nervous system and the maintenance
of nervous system architecture. Development (Cambridge, England), 132, 3691 -- 703. 129
[36] Child, C.M. (1911). Studies on the dynamics of morphogenesis and inheritance in ex-
perimental reproduction. I. The axial gradient in planaria dorotocephala as a limiting
factor in regulation. Journal of Experimental Zoology, 10, 265 -- 320. 4, 8
[37] Chong, T., Stary, J.M., Wang, Y. & Newmark, P.A. (2011). Molecular mark-
ers to characterize the hermaphroditic reproductive system of the planarian Schmidtea
mediterranea. BMC Developmental Biology, 11, 69. 10
152
REFERENCES
[38] Cochet-Escartin, O., Mickolajczyk, K.J. & Collins, E.M.S. (2015). Scrunching:
a novel escape gait in planarians. Physical Biology, 12, 056010. 59, 68
[39] Cross, M.C. & Hohenberg, P.C. (1993). Pattern formation outside of equilibrium.
Reviews of Modern Physics, 65, 851 -- 1112. 20
[40] Davidson, K.C., Adams, A.M., Goodson, J.M., McDonald, C.E., Potter, J.C.,
Berndt, J.D., Biechele, T.L., Taylor, R.J. & Moon, R.T. (2012). Wnt/β-catenin
signaling promotes differentiation, not self-renewal, of human embryonic stem cells and is
repressed by Oct4. Proceedings of the National Academy of Sciences of the United States
of America, 109, 4485 -- 90. 6
[41] Dobzhansky, T. (1973). Nothing in biology makes sense except in the light of evolution.
The American Biology Teacher , 35, 125 -- 129. 100
[42] Driesch, H. (1908). The science and philosophy of the organism. Adam and Charles
Black, London. 3
[43] Driever, W. & Nusslein-Volhard, C. (1988). A gradient of bicoid protein in
Drosophila embryos. Cell , 54, 83 -- 93. 5, 6
[44] Driever, W. & Nusslein-Volhard, C. (1988). The bicoid protein determines position
in the Drosophila embryo in a concentration-dependent manner. Cell , 54, 95 -- 104. 5, 23
[45] Driever, W., Siegel, V. & Nusslein-Volhard, C. (1990). Autonomous determi-
nation of anterior structures in the early Drosophila embryo by the bicoid morphogen.
Development (Cambridge, England), 109, 811 -- 820. 5
[46] Dubois, F. (1949). Contribution a l'´etude de la migration des cellules de r´eg´en´eration
chez les Planaires Dulcicoles. Ph.D. thesis, Strasbourg. 10
[47] Dubuis, J.O., Tkacik, G., Wieschaus, E.F., Gregor, T. & Bialek, W. (2013).
Positional information, in bits. Proceedings of the National Academy of Sciences, 110,
16301 -- 16308. 20
[48] Dunkel, J., Talbot, J. & Schotz, E.M. (2011). Memory and obesity affect the
population dynamics of asexual freshwater planarians. Physical biology, 8, 026003. 11,
129
[49] Economou, A.D., Ohazama, A., Porntaveetus, T., Sharpe, P.T., Kondo, S.,
Basson, M.A., Gritli-Linde, A., Cobourne, M.T. & Green, J.B.A. (2012). Peri-
odic stripe formation by a Turing mechanism operating at growth zones in the mammalian
palate. Nature Genetics, 44, 348 -- 351. 16
[50] Egger, B., Gschwentner, R. & Rieger, R. (2007). Free-living flatworms under the
knife: Past and present. Development Genes and Evolution, 217, 89 -- 104. 8, 9, 11, 57,
66, 134
153
REFERENCES
[51] Elliott, S.A. & S´anchez Alvarado, A. (2013). The history and enduring contribu-
tions of planarians to the study of animal regeneration. Wiley Interdisciplinary Reviews:
Developmental Biology, 2, 301 -- 326. 10, 54, 58
[52] Ephrussi, A. & St Johnston, D. (2004). Seeing Is Believing: The Bicoid Morphogen
Gradient Matures. Cell , 116, 143 -- 152. 5
[53] Flatt, T. (2012). A New Definition of Aging? Frontiers in Genetics, 3, 1 -- 2. 134
[54] Forsthoefel, D.J. & Newmark, P.A. (2009). Emerging patterns in planarian regen-
eration. Current opinion in genetics & development, 19, 412 -- 20. 12, 13, 20, 54
[55] Fried, P. & Iber, D. (2014). Dynamic scaling of morphogen gradients on growing
domains. Nature Communications, 5, 5077. 24, 98
[56] Frohnhofer, H.G. & Nusslein-Volhard, C. (1986). Organization of anterior pattern
in the Drosophila embryo by the maternal gene bicoid. Nature, 324, 120 -- 125. 5
[57] Fu, J., Posnien, N., Bolognesi, R., Fischer, T.D., Rayl, P., Oberhofer, G.,
Kitzmann, P., Brown, S.J. & Bucher, G. (2012). Asymmetrically expressed axin
required for anterior development in Tribolium. Proceedings of the National Academy of
Sciences, 109, 7782 -- 7786. 13
[58] Gallagher, T., Bjorness, T., Greene, R., You, Y.J. & Avery, L. (2013). The
geometry of locomotive behavioral states in C. elegans. PloS one, 8, e59865. 59
[59] Gao, B., Song, H., Bishop, K., Elliot, G., Garrett, L., English, M.A., Andre,
P., Robinson, J., Sood, R., Minami, Y., Economides, A.N. & Yang, Y. (2011).
Wnt Signaling Gradients Establish Planar Cell Polarity by Inducing Vangl2 Phosphory-
lation through Ror2. Developmental Cell , 20, 163 -- 176. 62
[60] Gao, C. & Chen, Y.G. (2010). Dishevelled: The hub of Wnt signaling. Cellular Sig-
nalling, 22, 717 -- 727. 62
[61] Gavino, M.A. & Reddien, P.W. (2011). A Bmp/Admp regulatory circuit controls
maintenance and regeneration of dorsal-ventral polarity in planarians. Current Biology,
21, 294 -- 299. 6, 57
[62] Gavino, M.A., Wenemoser, D., Wang, I.E. & Reddien, P.W. (2013). Tissue ab-
sence initiates regeneration through Follistatin-mediated inhibition of Activin signaling.
eLife, 2013, 1 -- 13. 13
[63] Gentile, L., Cebri`a, F. & Bartscherer, K. (2011). The planarian flatworm: an
in vivo model for stem cell biology and nervous system regeneration. Disease models &
mechanisms, 4, 12 -- 19. 8, 10, 11
[64] Gierer, A. (1981). Generation of biological patterns and form: Some physical, math-
ematical, and logical aspects. Prog. Biophys. Mol. Biol , 37, 1 -- 47. 15, 16, 17, 18, 38,
109
154
REFERENCES
[65] Gierer, A. (1988). Biological Pattern Formation and Physico-Chemical Laws. Synerget-
ics and Dynamic Instabilities, 151 -- 174. 15, 16, 17, 18, 109
[66] Gierer, A. & Meinhardt, H. (1972). A theory of biological pattern formation. Kyber-
netik , 12, 30 -- 9. 15, 19, 38
[67] Gierer, A. & Meinhardt, H. (1974). Biological pattern formation involving lateral
inhibition. In S.A. Levin, ed., Some Mathematical Questions in Biology. VI, Lectures on
Mathematics in the Life Sciences, Vol. 7 , 163 -- 183, American Mathematical Soc. 38
[68] Gilbert, S.F. (2014). Developmental Biology. Sinauer Associates, Incorporated, Sun-
derland, MA. 1, 2, 3, 4, 5, 7, 8, 9, 13, 20, 73
[69] Gonz´alez-Est´evez, C. (2009). Autophagy meets planarians. Autophagy, 5, 290 -- 297.
9, 10, 11, 12
[70] Gonz´alez-Est´evez, C., Felix, D.A., Aboobaker, A.A. & Sal´o, E. (2007). Gtdap-
1 promotes autophagy and is required for planarian remodeling during regeneration and
starvation. Proceedings of the National Academy of Sciences of the United States of Amer-
ica, 104, 13373 -- 13378. 11
[71] Gonz´alez-Est´evez, C., Felix, D.A., Rodr´ıguez-Esteban, G. & Aboobaker,
A.A. (2012). Decreased neoblast progeny and increased cell death during starvation-
induced planarian degrowth. The International journal of developmental biology, 56, 83 --
91. 77, 81, 87, 91, 97, 133, 137
[72] Green, J.B., New, H.V. & Smith, J.C. (1992). Responses of embryonic Xenopus cells
to activin and FGF are separated by multiple dose thresholds and correspond to distinct
axes of the mesoderm. Cell , 71, 731 -- 9. 5, 14
[73] Green, J.B., Smith, J.C. & Gerhart, J.C. (1994). Slow emergence of a multithresh-
old response to activin requires cell-contact-dependent sharpening but not prepattern.
Development (Cambridge, England), 120, 2271 -- 8. 14
[74] Green, J.B.A. & Sharpe, J. (2015). Positional information and reaction-diffusion:
two big ideas in developmental biology combine. Development (Cambridge, England),
142, 1203 -- 1211. 14, 15, 16, 42
[75] Green, J.B.A. & Smith, J.C. (1990). Graded changes in dose of a Xenopus activin A
homologue elicit stepwise transitions in embryonic cell fate. Nature, 347, 391 -- 394. 5, 14
[76] Gregor, T., Bialek, W., de Ruyter van Steveninck, R.R., Tank, D.W. & Wi-
eschaus, E.F. (2005). Diffusion and scaling during early embryonic pattern formation.
Proceedings of the National Academy of Sciences of the United States of America, 102,
18403 -- 7. 13, 20, 22, 23, 27, 57
[77] Gregor, T., Tank, D.W., Wieschaus, E.F. & Bialek, W. (2007). Probing the
limits to positional information. Cell , 130, 153 -- 64. 20
155
REFERENCES
[78] Gregor, T., Wieschaus, E.F., McGregor, A.P., Bialek, W. & Tank, D.W.
(2007). Stability and nuclear dynamics of the bicoid morphogen gradient. Cell , 130, 141 --
52. 23, 27, 28, 33
[79] Gurdon, J.B. & Bourillot, P.Y. (2001). Morphogen gradient interpretation. Nature,
413, 797 -- 803. 5
[80] Gurdon, J.B., Mitchell, A. & Mahony, D. (1995). Direct and continuous assessment
by cells of their position in a morphogen gradient. 5, 14
[81] Gurley, K., Elliott, S. & Simakov, O. (2010). Expression of secreted Wnt path-
way components reveals unexpected complexity of the planarian amputation response.
Developmental biology, 347, 24 -- 39. 12, 13, 54, 62, 98
[82] Gurley, K.A., Rink, J.C. & S´anchez Alvarado, A. (2008). Beta-catenin defines
head versus tail identity during planarian regeneration and homeostasis. Science (New
York, N.Y.), 319, 323 -- 7. 12, 54, 60, 61, 98
[83] Hamaratoglu, F., de Lachapelle, A.M., Pyrowolakis, G., Bergmann, S. &
Affolter, M. (2011). Dpp signaling activity requires Pentagone to scale with tissue size
in the growing Drosophila wing imaginal disc. PLoS biology, 9, e1001182. 24
[84] Handberg-Thorsager, M. & Sal´o, E. (2007). The planarian nanos-like gene Smednos
is expressed in germline and eye precursor cells during development and regeneration.
Development genes and evolution, 217, 403 -- 11. 10, 133
[85] Handberg-Thorsager, M., Fernandez, E. & Sal´o, E. (2008). Stem cells and re-
generation in planarians. Frontiers in bioscience : a journal and virtual library, 13, 6374 --
6394. 11
[86] Haranghy, L. & Bal´azs, A. (1964). Ageing and rejuvenation in planarians. Experi-
mental Gerontology, 1, 77 -- 91. 133, 134
[87] Haynie, D.T. (2001). Biological Thermodynamics. Cambridge University Press. 20
[88] Herpin, A. (2004). Transforming growth factor-β-related proteins: an ancestral and
widespread superfamily of cytokines in metazoans. Developmental & Comparative Im-
munology, 28, 461 -- 485. 5
[89] Howard, J., Grill, S.W. & Bois, J.S. (2011). Turing's next steps: the mechanochem-
ical basis of morphogenesis. Nature reviews. Molecular cell biology, 12, 392 -- 8. 13
[90] Hunding, A. & Sørensen, P.G. (1988). Size adaptation of Turing prepatterns. Journal
of mathematical biology, 26, 27 -- 39. 28, 33, 41, 42, 56
[91] Hyman, L.H. (1919). Physiological studies on Planaria. I. Oxygen consumption related
to feeding and starvation. American Journal of Physiology, 49, 377 -- 402. 134
[92] Hyman, L.H. (1919). Physiological Studies on Planaria. III . Oxygen Consumption in
Relation to Age (Size) Differences. Biological Bulletin, 37, 388 -- 403. 87, 89, 133
156
REFERENCES
[93] Hyman, L.H., Willier, B.H. & Rifenburgh, S.A. (1924). Physiological studies on
Planaria. VI. A respiratory and histochemical investigation of the source of the increased
metabolism after feeding. Journal of Experimental Zoology, 40, 473 -- 494. 90
[94] Inomata, H., Shibata, T., Haraguchi, T. & Sasai, Y. (2013). Scaling of Dorsal-
Ventral Patterning by Embryo Size-Dependent Degradation of Spemann's Organizer Sig-
nals. Cell , 153, 1296 -- 1311. 27
[95] Inoue, T., Kumamoto, H. & Okamoto, K. (2004). Morphological and functional re-
covery of the planarian photosensing system during head regeneration. Zoological science,
21, 275 -- 283. 9, 59
[96] Ishihara, S. & Kaneko, K. (2006). Turing pattern with proportion preservation. Jour-
nal of theoretical biology, 238, 683 -- 93. 24, 27, 28, 33, 38, 41, 42, 56
[97] Jackson, J.E. (2005). A user's guide to principal components. John Wiley & Sons. 59
[98] Jenkins, M.M. (1963). Bipolar Planarians in a Stock Culture. Science (New York, N.Y.),
142, 1187. 57
[99] Jolliffe, I. (2005). Principal component analysis. Wiley Online Library. 59
[100] Kalthoff, K. (1976). Analysis of a Cytoplasmic Determinant in an Insect Egg. In
M. Karkinen-Jaaskelainen & L. Weiss, eds., Cell Interactions in differentiation, 5 -- 21,
Academic Press, London. 5
[101] Kang, H. & Alvarado, A.S. (2009). Flow cytometry methods for the study of cell-cycle
parameters of planarian stem cells. Developmental Dynamics, 238, 1111 -- 1117. 91
[102] Kato, K., Orii, H., Watanabe, K. & Agata, K. (1999). The role of dorsoventral
interaction in the onset of planarian regeneration. Development (Cambridge, England),
126, 1031 -- 1040. 56
[103] Kavanagh, K.D., Shoval, O., Winslow, B.B., Alon, U., Leary, B.P., Kan, A.
& Tabin, C.J. (2013). Developmental bias in the evolution of phalanges. Proceedings of
the National Academy of Sciences, 110, 18190 -- 18195. 59, 69
[104] Kiecker, C. & Niehrs, C. (2001). A morphogen gradient of Wnt/beta-catenin sig-
nalling regulates anteroposterior neural patterning in Xenopus. Development (Cambridge,
England), 128, 4189 -- 4201. 14
[105] Kirk, G.S. (1954). Heraclitus: The Cosmic Fragments. Cambridge University Press. 70
[106] Koch, A. & Meinhardt, H. (1994). Biological pattern formation: from basic mecha-
nisms to complex structures. Reviews of Modern Physics, 66. 15
[107] Komiya, Y. & Habas, R. (2008). Wnt signal transduction pathways. Organogenesis, 4,
68 -- 75. 6, 60, 62
[108] Kriss, T.C. & Kriss, V.M. (1998). History of the Operating Microscope: From Mag-
nifying Glass to Microneurosurgery. Neurosurgery, 42, 899 -- 907. 3
157
REFERENCES
[109] Kruse, K., Pantazis, P., Bollenbach, T., Julicher, F. & Gonz´alez-Gait´an, M.
(2004). Dpp gradient formation by dynamin-dependent endocytosis: receptor trafficking
and the diffusion model. Development (Cambridge, England), 131, 4843 -- 56. 5, 6
[110] Lange, C.S. (1967). A quantitative study of the number and distribution of neoblasts in
Dugesia lugubris (Planaria) with reference to size and ploidy. Journal of embryology and
experimental morphology, 18, 199 -- 213. 75, 76, 83
[111] Lange, C.S. (1968). A possible explanation in cellular terms of the physiological ageing
of the planarian. Experimental gerontology, 3, 219 -- 230. 75, 76, 83, 97, 133
[112] Lawrence, P.A. (2001). Morphogens: how big is the big picture? Nature cell biology,
3, E151 -- E154. 13
[113] Layalle, S., Arquier, N. & L´eopold, P. (2008). The TOR Pathway Couples Nu-
trition and Developmental Timing in Drosophila. Developmental Cell , 15, 568 -- 577. 20,
70
[114] LeRoith, D., Shiloach, J., Heffron, R., Rubinovitz, C., Tanenbaum, R. &
Roth, J. (1985). Insulin-related material in microbes: similarities and differences from
mammalian insulins. Canadian journal of biochemistry and cell biology = Revue canadi-
enne de biochimie et biologie cellulaire, 63, 839 -- 49. 87
[115] Levin, M. (2012). Molecular bioelectricity in developmental biology: new tools and recent
discoveries: control of cell behavior and pattern formation by transmembrane potential
gradients. BioEssays: news and reviews in molecular, cellular and developmental biology,
34, 205 -- 17. 57
[116] Liu, S.Y., Selck, C., Friedrich, B., Lutz, R., Vila-Farr´e, M., Dahl, A.,
Brandl, H., Lakshmanaperumal, N., Henry, I. & Rink, J.C. (2013). Reactivating
head regrowth in a regeneration-deficient planarian species. Nature, 500, 81 -- 4. 12, 66
[117] Lobo, D., Beane, W.S. & Levin, M. (2012). Modeling planarian regeneration: a
primer for reverse-engineering the worm. PLoS computational biology, 8, e1002481. 12,
13, 54, 98, 99
[118] Ma, R., Klindt, G.S., Riedel-Kruse, I.H., Julicher, F. & Friedrich, B.M.
(2014). Active Phase and Amplitude Fluctuations of Flagellar Beating. Physical Review
Letters, 113, 048101. 59
[119] Marcon, L. & Sharpe, J. (2012). Turing patterns in development: what about the
horse part? Current opinion in genetics & development, 22, 578 -- 84. 16
[120] Meinhardt, H. (1982). Models of biological pattern formation. Academic Press, London.
15, 16, 18, 20, 38
[121] Meinhardt, H. (2004). Different strategies for midline formation in bilaterians. Nature
reviews. Neuroscience, 5, 502 -- 10. 99
158
REFERENCES
[122] Meinhardt, H. (2009). Models for the generation and interpretation of gradients. Cold
Spring Harbor perspectives in biology, 1, a001362. 57
[123] Michael Bate & Alfonso Martinez Arias, eds. (1993). The Development of
Drosophila Melanogaster , vol. 1. Cold Spring Harbor Laboratory Press, New York. 24
[124] Milo, R., Jorgensen, P., Moran, U., Weber, G. & Springer, M. (2010).
BioNumbers -- the database of key numbers in molecular and cell biology. Nucleic acids
research, 38, D750 -- 3. 70, 84, 90
[125] Mocek, R. (1971). Wilhelm Roux und Hans Driesch - Zur Geschichte der Entwick-
lungsphysiologie der Tiere. Gustav Fischer, Jena. 3
[126] Montgomery, J.R. & Coward, S.J. (1974). On the minimal size of a planarian capable
of regeneration. Transactions of the American Microscopical Society, 93, 386 -- 391. 9
[127] Morgan, T.H. (1898). Experimental studies of the regeneration of Planaria maculata.
Archiv fur Entwicklungsmechanik der Organismen, 7, 364 -- 397. 54
[128] Morgan, T.H. (1901). Regeneration. New York, The Macmillan Company; London,
Macmillan & Co., ltd. 3, 4, 8, 9, 12
[129] Morgan, T.H. (1905). "Polarity" considered as a phenomenon of gradation of materials.
Journal of Experimental Zoology, 2, 495 -- 506. 4, 8
[130] Morgan, T.H. (1927). Experimental embryology. Columbia University Press, New York.
22
[131] Morita, M. & Best, J.B. (1984). Effects of photoperiods and melatonin on planarian
asexual reproduction. Journal of Experimental Zoology, 231, 273 -- 282. 11
[132] Moritz, S., Stockle, F., Ortmeier, C., Schmitz, H., Rodr´ıguez-Esteban, G.,
Key, G. & Gentile, L. (2012). Heterogeneity of planarian stem cells in the S/G2/M
phase. The International journal of developmental biology, 56, 117 -- 25. 11
[133] Mouton, S., Willems, M., Houthoofd, W., Bert, W. & Braeckman, B.P.
(2011). Lack of metabolic ageing in the long-lived flatworm Schmidtea polychroa. Ex-
perimental Gerontology, 46, 755 -- 761. 75, 76, 84, 87, 89, 90, 133
[134] Muller, P., Rogers, K.W., Jordan, B.M., Lee, J.S., Robson, D., Ramanathan,
S. & Schier, a.F. (2012). Differential Diffusivity of Nodal and Lefty Underlies a
Reaction-Diffusion Patterning System. Science, 336, 721 -- 724. 16
[135] Mumcu, P. (2011). Self-organized Growth in Developing Epithelia. Ph.D. thesis, TU
Dresden. 14, 24, 27, 28, 29, 30, 31, 33, 41
[136] Nakamasu, A., Takahashi, G., Kanbe, A. & Kondo, S. (2009). Interactions between
zebrafish pigment cells. Proceedings of the National Academy of Sciences, 106, 8429 -- 8434.
16
159
REFERENCES
[137] Nakamura, T., Mine, N., Nakaguchi, E., Mochizuki, A., Yamamoto, M.,
Yashiro, K., Meno, C. & Hamada, H. (2006). Generation of Robust Left-Right Asym-
metry in the Mouse Embryo Requires a Self-Enhancement and Lateral-Inhibition System.
Developmental Cell , 11, 495 -- 504. 16
[138] Neumann, C. & Cohen, S. (1997). Morphogens and pattern formation. BioEssays, 19,
721 -- 729. 5
[139] Newmark, P.A. & S´anchez Alvarado, A. (2000). Bromodeoxyuridine specifically
labels the regenerative stem cells of planarians. Developmental biology, 220, 142 -- 153. 77,
91, 138
[140] Newmark, P.A. & S´anchez Alvarado, A. (2002). Not your father's planarian: a
classic model enters the era of functional genomics. Nature reviews. Genetics, 3, 210 -- 9.
8, 9, 10, 11, 57, 71, 133, 134
[141] Niehrs, C. (2010). On growth and form: a Cartesian coordinate system of Wnt and
BMP signaling specifies bilaterian body axes. Development (Cambridge, England), 137,
845 -- 857. 3, 6, 12, 54, 57
[142] Nimeth, K.T., Mahlknecht, M., Mezzanato, A., Peter, R., Rieger, R. &
Ladurner, P. (2004). Stem Cell Dynamics during Growth, Feeding, and Starvation
in the Basal Flatworm Macrostomum sp. (Platyhelminthes). Developmental Dynamics,
230, 91 -- 99. 77, 91, 138
[143] Nusse, R., He, X. & van Amerongen, R., eds. (2013). Wnt Signaling: A Subject Col-
lection from Cold Spring Harbor Perspectives in Biology. Cold Spring Harbor Laboratory
Press, New York. 7, 9
[144] Oberhofer, G., Grossmann, D., Siemanowski, J.L., Beissbarth, T. & Bucher,
G. (2014). Wnt/β-catenin signaling integrates patterning and metabolism of the insect
growth zone. Development (Cambridge, England), 1 -- 11. 6
[145] Okubo, A. (1980). Diffusion and ecological problems: Mathematical models. Springer,
Berlin. 19
[146] Okubo, A. & Levin, S.A. (2002). Diffusion and Ecological Problems, Modern Perspec-
tives, Second Edition. Springer. 17, 18, 19, 109
[147] Oster, G.F. (1988). Lateral inhibition models of developmental processes. Mathematical
Biosciences, 90, 265 -- 286. 13, 16, 18, 55
[148] Othmer, H.G. & Pate, E. (1980). Scale-invariance in reaction-diffusion models of spa-
tial pattern formation. Proceedings of the National Academy of Sciences of the United
States of America, 77, 4180 -- 4. 24, 27, 28, 33, 38, 41, 42, 56
[149] Oviedo, N.J. & Levin, M. (2007). Smedinx-11 Is a Planarian Stem Cell Gap Junction
Gene Required for Regeneration and Homeostasis. Development (Cambridge, England),
134, 3121 -- 3131. 13
160
REFERENCES
[150] Oviedo, N.J., Newmark, P.A. & S´anchez Alvarado, A. (2003). Allometric scaling
and proportion regulation in the freshwater planarian Schmidtea mediterranea. Develop-
mental Dynamics, 226, 326 -- 333. 71, 75, 76, 81, 134, 135, 136
[151] Oviedo, N.J., Nicolas, C.L., Adams, D.S. & Levin, M. (2008). Planarians: a ver-
satile and powerful model system for molecular studies of regeneration, adult stem cell
regulation, aging, and behavior. CSH protocols, 2008, pdb.emo101. 8, 11, 71
[152] Oviedo, N.J., Morokuma, J., Walentek, P., Kema, I.P., Gu, M.B., Ahn, J.M.,
Hwang, J.S., Gojobori, T. & Levin, M. (2010). Long-range neural and gap junc-
tion protein-mediated cues control polarity during planarian regeneration. Developmental
Biology, 339, 188 -- 199. 9, 55, 57
[153] Paskin, T.R., Jellies, J., Bacher, J. & Beane, W.S. (2014). Planarian Phototac-
tic Assay Reveals Differential Behavioral Responses Based on Wavelength. PloS one, 9,
e114708. 9
[154] Pate, E. & Othmer, H.G. (1984). Applications of a model for scale-invariant pattern
formation in developing systems. Differentiation, 28, 1 -- 8. 24, 27, 28, 33, 41, 42, 56
[155] Pavlova, G.A. (2000). Ciliary locomotion of the mollusk is different from that of the
planaria. Doklady biological sciences : proceedings of the Academy of Sciences of the
USSR, Biological sciences sections / translated from Russian, 375, 630 -- 2. 60
[156] Pearson, K. (1901). On lines and planes of closest fit to systems of points in space.
Philosophical Magazine Series 6 , 2, 559 -- 572. 59
[157] Peiris, T.H. & Oviedo, N.J. (2013). Gap junction proteins: Master regulators of the
planarian stem cell response to tissue maintenance and injury. Biochimica et Biophysica
Acta - Biomembranes, 1828, 109 -- 117. 57
[158] Pellettieri, J. & S´anchez Alvarado, A. (2007). Cell turnover and adult tissue
homeostasis: from humans to planarians. Annual review of genetics, 41, 83 -- 105. 1, 8, 9,
10, 11, 12, 20, 70, 71, 134
[159] Pellettieri, J., Fitzgerald, P., Watanabe, S., Mancuso, J., Green, D.R. &
S´anchez Alvarado, A. (2010). Cell death and tissue remodeling in planarian regener-
ation. Developmental biology, 338, 76 -- 85. 10, 71, 91
[160] Petersen, C.P. & Reddien, P.W. (2008). Smed-betacatenin-1 is required for antero-
posterior blastema polarity in planarian regeneration. Science (New York, N.Y.), 319,
327 -- 330. 12, 54
[161] Pigon, A., Morita, M. & Best, J.B. (1974). Cephalic mechanism for social control
of fissioning in planarians. II. Localization and identification of the receptors by election
micrographic and ablation studies. Journal of Neurobiology, 5, 443 -- 462. 11, 129
[162] Prados, J., Alvarez, B., Howarth, J., Stewart, K., Gibson, C.L., Hutchinson,
C.V., Young, A.M.J. & Davidson, C. (2013). Cue competition effects in the planarian.
Animal Cognition, 16, 177 -- 186. 8
161
REFERENCES
[163] Press, W., Teukolsky, S., Vetterling, W., Flannery, B., Ziegel, E., Press,
W., Flannery, B., Teukolsky, S. & Vetterling, W. (2002). Numerical Recipes:
The Art of Scientific Computing, vol. 29. Cambridge University Press, 2nd edn. 125
[164] Raspopovic, J., Marcon, L., Russo, L. & Sharpe, J. (2014). Digit patterning
is controlled by a Bmp-Sox9-Wnt Turing network modulated by morphogen gradients.
Science, 566, 10 -- 15. 6, 16
[165] Reddien, P.W. & S´anchez Alvarado, A. (2004). Fundamentals of planarian regen-
eration. Annual review of cell and developmental biology, 20, 725 -- 57. 9, 10, 12
[166] Reed, J.C. (1999). Dysregulation of apoptosis in cancer. Journal of Clinical Oncology,
17, 2941 -- 2953. 70
[167] Richards, D.M. & Saunders, T.E. (2015). Spatiotemporal analysis of different mech-
anisms for interpreting morphogen gradients. Biophysical journal , 108, 2061 -- 73. 14, 24
[168] Ringelnatz, J. (1924). Abseits der Geographie. In Nervosipopel. Elf Angelegenheiten,
chap. 2, Gunther Langes, Munchen. 11
[169] Rink, J., Gurley, K., Elliott, S. & S´anchez Alvarado, A. (2009). Planarian Hh
signaling regulates regeneration polarity and links Hh pathway evolution to cilia. Science,
326, 1406. 9, 57, 58, 59, 66
[170] Rink, J.C. (2013). Stem cell systems and regeneration in planaria. Development Genes
and Evolution, 223, 67 -- 84. 8, 9, 10, 11, 12, 71, 91, 138
[171] Rink, J.C., Vu, H.T.K. & S´anchez Alvarado, A. (2011). The maintenance and
regeneration of the planarian excretory system are regulated by EGFR signaling. Devel-
opment (Cambridge, England), 138, 3769 -- 80. 9, 10
[172] Rivera-Pomar, R. & Jackle, H. (1996). From gradients to stripes in Drosophila
embryogenesis: filling in the gaps. Trends in Genetics, 12, 478 -- 483. 13
[173] Roberts-Galbraith, R.H. & Newmark, P.A. (2015). On the organ trail:
insights
into organ regeneration in the planarian. Current Opinion in Genetics & Development,
32, 37 -- 46. 10, 68
[174] Rogers, K.W. & Schier, A.F. (2011). Morphogen gradients: from generation to in-
terpretation. Annual review of cell and developmental biology, 27, 377 -- 407. 4, 5, 13
[175] Romanczuk, P. & Salbreux, G. (2015). Optimal chemotaxis in intermittent migration
of animal cells. Physical Review E , 91, 42720. 14
[176] Romero, R. & Bagun`a, J. (1988). Quantitative cellular analysis of cell cycle strategies
of iteroparous and semelparous triclads. Fortschritte der Zoologie, 36, 283 -- 289. 137
[177] Romero, R. & Bagun`a, J. (1991). Quantitative cellular analysis of growth and repro-
duction in freshwater planarians (Turbellaria; Tricladida). I. A cellular description of the
intact organism. 76
162
REFERENCES
[178] Romero, R., Bagun`a, J. & Calow, P. (1991). Intraspecific Variation in Somatic Cell
Turnover and Regenerative Rate in the Freshwater Planarian Dendrocoelum Lacteum.
Invertebrate Reproduction & Development, 20, 107 -- 113. 137
[179] Rompolas, P., Patel-King, R.S. & King, S.M. (2009). Schmidtea mediterranea: a
model system for analysis of motile cilia., vol. 93. Elsevier, first edit edn. 8, 9, 58, 59, 66
[180] Rompolas, P., Patel-King, R.S. & King, S.M. (2010). An outer arm Dynein confor-
mational switch is required for metachronal synchrony of motile cilia in planaria. Molec-
ular biology of the cell , 21, 3669 -- 79. 9, 58, 59, 66
[181] Ruan, S. (1998). Diffusion-driven instability in the Gierer-Meinhardt model of morpho-
genesis. Natural Resource Modeling, 11. 19
[182] Sal´o, E. & Bagun`a, J. (1984). Regeneration and pattern formation in planarians. J.
Embryol exp. Morph., 83, 63 -- 80. 138
[183] Sal´o, E., Abril, J.F., Adell, T., Cebri`a, F., Eckelt, K., Fernandez-Taboada,
E., Handberg-Thorsager, M., Iglesias, M., Molina, M.D. & Rodr´ıguez-
Esteban, G. (2009). Planarian regeneration: achievements and future directions after
20 years of research. The International journal of developmental biology, 53, 1317 -- 1327.
8, 9, 11
[184] S´anchez Alvarado, A. & Tsonis, P.a. (2006). Bridging the regeneration gap: genetic
insights from diverse animal models. Nature reviews. Genetics, 7, 873 -- 884. 1, 8, 9, 68
[185] S´anchez Alvarado, A., Newmark, P.A., Robb, S.M. & Juste, R. (2002). The
Schmidtea mediterranea database as a molecular resource for studying platyhelminthes,
stem cells and regeneration. Development (Cambridge, England), 129, 5659 -- 65. 129
[186] Sander, K. (1960). Analyse des ooplasmatischen Reaktionssystems vonEuscelis plebe-
jus Fall (Cicadina) durch Isolieren und Kombinieren von Keimteilen II. Mitteilung Die
Differenzierungsleistungen nach Verlagern von Hinterpolmaterial. Wilhelm Roux' Archiv
fur Entwicklungsmechanik der Organismen, 151, 660 -- 707. 5
[187] Sander, K. (1975). Pattern Specification in the Insect Embryo. In R. Porter & J. Rivers,
eds., Ciba Foundation Symposium 29 - Cell Patterning, chap. 12, 241 -- 263, John Wiley
& Sons, Ltd., Chichester, UK. 5
[188] Sander, K. (1997). Landmarks in Developmental Biology 1883-1924 , vol. 53. Springer,
Berlin Heidelberg. 3, 4
[189] Sanger, T.D. (2000). Human arm movements described by a low-dimensional superpo-
sition of principal components. The Journal of neuroscience : the official journal of the
Society for Neuroscience, 20, 1066 -- 72. 59
[190] Savage, V.M., Allen, A.P., Brown, J.H., Gillooly, J.F., Herman, A.B.,
Woodruff, W.H. & West, G.B. (2007). Scaling of number, size, and metabolic rate
of cells with body size in mammals. Proceedings of the National Academy of Sciences of
the United States of America, 104, 4718 -- 4723. 87, 89
163
REFERENCES
[191] Schier, A.F. (2009). Nodal morphogens. Cold Spring Harbor Perspectives in Biology, 1,
a003459 -- a003459. 14, 16
[192] Schroedinger, E. (1944). What is life? . Cambridge University Press. 1
[193] Schultz, E. (1904). Uber Reduktionen. I. Uber Hungerserscheinungen bei Planaria
lactea. Archiv fur Entwicklungsmechanik der Organismen, 18, 555 -- 577. 71, 133
[194] Scimone, M.L., Kravarik, K.M., Lapan, S.W. & Reddien, P.W. (2014). Neoblast
Specialization in Regeneration of the Planarian Schmidtea mediterranea. Stem Cell Re-
ports, 3, 1 -- 14. 11
[195] Segel, L.A. & Jackson, J.L. (1972). Dissipative structure: an explanation and an
ecological example. Journal of theoretical biology, 37, 545 -- 59. 16, 17, 109
[196] Seifert, J.R.K. & Mlodzik, M. (2007). Frizzled/PCP signalling: a conserved mecha-
nism regulating cell polarity and directed motility. Nature Reviews Genetics, 8, 126 -- 138.
58
[197] Sheth, R., Marcon, L., Bastida, M.F., Junco, M., Quintana, L., Dahn, R.,
Kmita, M., Sharpe, J. & Ros, M.A. (2012). Hox Genes Regulate Digit Patterning by
Controlling the Wavelength of a Turing-Type Mechanism. Science, 338, 1476 -- 1480. 16
[198] Shomrat, T. & Levin, M. (2013). An automated training paradigm reveals long-term
memory in planaria and its persistence through head regeneration. The Journal of exper-
imental biology. 8
[199] Shoval, O., Sheftel, H., Shinar, G., Hart, Y., Ramote, O., Mayo, A., Dekel,
E., Kavanagh, K. & Alon, U. (2012). Evolutionary trade-offs, Pareto optimality, and
the geometry of phenotype space. Science (New York, N.Y.), 336, 1157 -- 60. 59, 69
[200] Sikes, J.M. & Newmark, P.A. (2013). Restoration of anterior regeneration in a pla-
narian with limited regenerative ability. Nature, 500, 77 -- 80. 12
[201] Simanov, D., Mellaart-Straver, I., Sormacheva, I. & Berezikov, E. (2012).
The flatworm macrostomum lignano is a powerful model organism for ion channel and
stem cell research. Stem Cells International , 2012. 13
[202] Sluys, R., Timoshkin, O.A. & Kawakatsu, M. (1998). A new species of giant pla-
narian from Lake Baikal, with some remarks on character states in the Dendrocoelidae
(Platyhelminthes, Tricladida, Paludicola). Hydrobiologia, 383, 69 -- 75. 8
[203] Spalding, K.L., Bergmann, O., Alkass, K., Bernard, S., Salehpour, M., Hut-
tner, H.B., Bostrom, E., Westerlund, I., Vial, C., Buchholz, B.A., Possnert,
G., Mash, D.C., Druid, H. & Fris´en, J. (2013). Dynamics of hippocampal neuroge-
nesis in adult humans. Cell , 153, 1219 -- 27. 92, 138, 139
[204] Spemann, H. & Mangold, H. (1924). Uber Induktion von Embryonalanlagen durch
Implantation artfremder Organisatoren. Roux's Arch. Entw. Mech., 599 -- 638. 5
164
REFERENCES
[205] Spencer, F.a., Hoffmann, F.M. & Gelbart, W.M. (1982). Decapentaplegic: a gene
complex affecting morphogenesis in Drosophila melanogaster. Cell , 28, 451 -- 461. 5
[206] St Johnston, D. & Nusslein-Volhard, C. (1992). The origin of pattern and polarity
in the Drosophila embryo. Cell , 68, 201 -- 19. 24
[207] Stephens, G.J., Johnson-Kerner, B., Bialek, W. & Ryu, W.S. (2008). Dimen-
sionality and dynamics in the behavior of C. elegans. PLoS Comp Biol , 4, e1000028. 59,
99
[208] Stephens, G.J., Johnson-Kerner, B., Bialek, W. & Ryu, W.S. (2010). From
modes to movement in the behavior of Caenorhabditis elegans. PloS one, 5, e13914. 59,
68, 99
[209] Stephens, G.J., Bueno de Mesquita, M., Ryu, W.S. & Bialek, W. (2011). Emer-
gence of long timescales and stereotyped behaviors in Caenorhabditis elegans. Proceedings
of the National Academy of Sciences, 108, 7286 -- 7289. 59, 99
[210] Sullivan, L.H. (1896). The Tall Office Building Artistically Considered. Lippincott's
Magazine, 403 -- 409. 58
[211] Takeda, H., Nishimura, K. & Agata, K. (2009). Planarians maintain a constant ratio
of different cell types during changes in body size by using the stem cell system. Zoological
science, 26, 805 -- 813. 75, 76
[212] Talbot, J. & Schotz, E.M. (2011). Quantitative characterization of planarian wild-
type behavior as a platform for screening locomotion phenotypes. The Journal of experi-
mental biology, 214, 1063 -- 1067. 8, 60
[213] Tan, T.C.J., Rahman, R., Jaber-Hijazi, F., Felix, D.A., Chen, C., Louis, E.J. &
Aboobaker, A. (2012). Telomere maintenance and telomerase activity are differentially
regulated in asexual and sexual worms. Proceedings of the National Academy of Sciences,
109, 4209 -- 4214. 133
[214] Tata, P.R., Mou, H., Pardo-Saganta, A., Zhao, R., Prabhu, M., Law, B.M.,
Vinarsky, V., Cho, J.L., Breton, S., Sahay, A., Medoff, B.D. & Rajagopal,
J. (2013). Dedifferentiation of committed epithelial cells into stem cells in vivo. Nature,
503, 218 -- 223. 3
[215] Thomas, M.A., Quinodoz, S. & Schotz, E.M. (2012). Size Matters! Journal of
Statistical Physics, 148, 664 -- 676. 81, 133
[216] Thompson, D.W. (1910). A history of animals. Clarendon Press, Oxford. 2
[217] Thompson, D.W. (1945). On Growth and Form. Cambridge : University Press ; New
York : Macmillan, 2nd edn. 1, 42
[218] Tkacik, G. & Bialek, W. (2014). Information processing in living systems. arXiv ,
1 -- 21. 20
165
REFERENCES
[219] Tkacik, G. & Walczak, A.M. (2011). Information transmission in genetic regulatory
networks: a review. Journal of physics. Condensed matter , 23, 153102. 20
[220] Travis, G. (1981). Replicating Replication? Aspects of the Social Construction of Learn-
ing in Planarian Worms. Social Studies of Science, 11, 11 -- 32. 9
[221] Turing, A.M. (1952). The Chemical Basis of Morphogenesis. Philosophical Transactions
of the Royal Society of London, Series B , 237, 37 -- 72. 5, 13, 15, 16, 17, 109
[222] Umesono, Y., Tasaki, J., Nishimura, Y., Hrouda, M., Kawaguchi, E., Yazawa,
S., Nishimura, O., Hosoda, K., Inoue, T. & Agata, K. (2013). The molecular logic
for planarian regeneration along the anterior-posterior axis. Nature, 500, 73 -- 6. 12
[223] Umulis, D.M. & Othmer, H.G. (2013). Mechanisms of scaling in pattern formation.
Development, 140, 4830 -- 43. 13, 20, 22, 24, 26, 27, 42
[224] Urdy, S. (2012). On the evolution of morphogenetic models: mechano-chemical inter-
actions and an integrated view of cell differentiation, growth, pattern formation and
morphogenesis. Biological Reviews, 87, 786 -- 803. 13
[225] van Wolfswinkel, J.C., Wagner, D.E. & Reddien, P.W. (2014). Single-Cell Anal-
ysis Reveals Functionally Distinct Classes within the Planarian Stem Cell Compartment.
Cell Stem Cell , 1 -- 14. 11
[226] Vladar, E.K., Antic, D. & Axelrod, J.D. (2009). Planar cell polarity signaling: the
developing cell's compass. Cold Spring Harbor perspectives in biology, 1, a002964. 58
[227] von Bertalanffy, L. (1940). Untersuchungen uber die Gesetzlichkeit des Wachstums -
III. Teil. Quantitative Beziehungen zwischen Darmoberflache und Korpergrosse bei Pla-
naria maculata. Wilhelm Roux' Archiv fur Entwicklungsmechanik der Organismen, 140,
81 -- 89. 75, 76
[228] von Goethe, J.W. (1933). Faust - Der Tragodie zweiter Teil . J.G. Cotta'sche Buch-
handlung, Stuttgart, Tubingen. 3
[229] Wagner, D.E., Wang, I.E. & Reddien, P.W. (2011). Clonogenic neoblasts are
pluripotent adult stem cells that underlie planarian regeneration. Science (New York,
N.Y.), 332, 811 -- 6. 10, 134
[230] Wang, Y., Zayas, R.M., Guo, T. & Newmark, P.A. (2007). Nanos Function Is
Essential for Development and Regeneration of Planarian Germ Cells. Proceedings of the
National Academy of Sciences of the United States of America, 104, 5901 -- 5906. 133
[231] Wang, Y., Stary, J.M., Wilhelm, J.E. & Newmark, P.A. (2010). A functional
genomic screen in planarians identifies novel regulators of germ cell development. Genes
& development, 24, 2081 -- 92. 10
[232] Wartlick, O., Kicheva, A. & Gonz´alez-Gait´an, M. (2009). Morphogen gradient
formation. Cold Spring Harbor perspectives in biology, 1, a001255. 13, 20, 23
166
REFERENCES
[233] Wartlick, O., Mumcu, P., Julicher, F. & Gonz´alez-Gait´an, M. (2011). Under-
standing morphogenetic growth control -- lessons from flies. Nature reviews. Molecular cell
biology, 12, 594 -- 604. 14, 24, 27, 28, 33, 41, 98
[234] Wartlick, O., Mumcu, P., Kicheva, A., Bittig, T., Seum, C., Julicher, F. &
Gonz´alez-Gait´an, M. (2011). Dynamics of Dpp signaling and proliferation control.
Science (New York, N.Y.), 331, 1154 -- 9. 5, 14, 23, 24, 27, 28, 33, 41, 42
[235] Wartlick, O., Julicher, F. & Gonz´alez-Gait´an, M. (2014). Growth control by
a moving morphogen gradient during Drosophila eye development. Development, 141,
1884 -- 93. 14, 24
[236] Watanabe, M. & Kondo, S. (2015). Is pigment patterning in fish skin determined by
the Turing mechanism? Trends in Genetics, 31, 88 -- 96. 16
[237] Wenemoser, D. & Reddien, P.W. (2010). Planarian regeneration involves distinct
stem cell responses to wounds and tissue absence. Developmental Biology, 344, 979 -- 991.
12
[238] Werner, S., Rink, J.C., Riedel-Kruse, I.H. & Friedrich, B.M. (2014). Shape
Mode Analysis Exposes Movement Patterns in Biology: Flagella and Flatworms as Case
Studies. PLoS ONE , 9, e113083. 58
[239] Werner, S., Stuckemann, T., Beir´an Amigo, M., Rink, J.C., Julicher, F. &
Friedrich, B.M. (2015). Scaling and Regeneration of Self-Organized Patterns. Physical
Review Letters, 114. 22, 42
[240] West, G.B., Woodruff, W.H. & Brown, J.H. (2002). Allometric scaling of
metabolic rate from molecules and mitochondria to cells and mammals. Proceedings of the
National Academy of Sciences of the United States of America, 99 Suppl 1, 2473 -- 2478.
84, 90
[241] Whittaker, J.R. (1982). Muscle lineage cytoplasm can change the developmental ex-
pression in epidermal lineage cells of ascidian embryos. Dev Biol , 93, 463 -- 470. 3
[242] Wilson, P.a. & Melton, D.a. (1994). Mesodermal patterning by an inducer gradient
depends on secondary cell-cell communication. Curr Biol , 4, 676 -- 86. 14
[243] Wolpert, L. (1969). Positional information and the spatial pattern of cellular differen-
tiation. Journal of theoretical biology, 25, 1 -- 47. 5, 13
[244] Wolpert, L. & Tickle, C. (2011). Principles of Development. Oxford University Press,
New York. 1, 2, 3, 5, 6, 7, 8, 9, 13, 20, 68, 70
[245] Yazawa, S., Umesono, Y., Hayashi, T., Tarui, H. & Agata, K. (2009). Planarian
Hedgehog/Patched establishes anterior-posterior polarity by regulating Wnt signaling.
Proceedings of the National Academy of Sciences of the United States of America, 106,
22329 -- 34. 10, 54, 57
167
Eigenstandigkeitserklarung
Hiermit versichere ich, dass ich die vorliegende Arbeit ohne unzulassige Hilfe
Dritter und ohne Benutzung anderer als der angegebenen Hilfsmittel ange-
fertigt habe; die aus fremden Quellen direkt oder indirekt ubernommenen
Gedanken sind als solche kenntlich gemacht. Die Arbeit wurde bisher weder
im Inland noch im Ausland in gleicher oder ahnlicher Form einer anderen
Prufungsbehorde vorgelegt.
Die vorliegende Arbeit wurde im Zeitraum zwischen Oktober 2011 und
Dezember 2015 unter der Betreuung durch Prof. Dr. Frank Julicher und
Dr. Benjamin M. Friedrich am Max-Planck-Institut fur Physik komplexer
Systeme in Dresden angefertigt.
Ich versichere, dass ich bisher keine erfolglosen Promotionsverfahren un-
ternommen habe.
Ich erkenne die Promotionsordnung der Fakultat der
Mathematik und Naturwissenschaften der Technischen Universitat Dresden
an.
Teile dieser Arbeit wurden schon in folgender Form veroffentlicht:
- S. Werner, J.C. Rink, I.-H. Riedel-Kruse, B.M. Friedrich: Shape mode
analysis exposes movement patterns in biology, PLoS ONE, 9(11), e113083,
2014
- S. Werner, T. Stuckemann, M. Beiran Amigo, J.C. Rink, F. Julicher, B.M.
Friedrich: Scaling and regeneration of self-organized patterns, Phys. Rev.
Lett. 114, 138101, 2015 (Editors' suggestion)
Dresden,
|
1007.0315 | 1 | 1007 | 2010-07-02T08:08:52 | A simple theory of protein folding kinetics | [
"physics.bio-ph",
"physics.chem-ph",
"q-bio.BM"
] | We present a simple model of protein folding dynamics that captures key qualitative elements recently seen in all-atom simulations. The goals of this theory are to serve as a simple formalism for gaining deeper insight into the physical properties seen in detailed simulations as well as to serve as a model to easily compare why these simulations suggest a different kinetic mechanism than previous simple models. Specifically, we find that non-native contacts play a key role in determining the mechanism, which can shift dramatically as the energetic strength of non-native interactions is changed. For protein-like non-native interactions, our model finds that the native state is a kinetic hub, connecting the strength of relevant interactions directly to the nature of folding kinetics. | physics.bio-ph | physics |
A simple theory of protein folding kinetics
Departments of Chemistry, Structural Biology, and Computer Science, Stanford University, Stanford, CA
(Dated: May 29, 2018)
Vijay S. Pande
We present a simple model of protein folding dynamics that captures key qualitative elements
recently seen in all-atom simulations. The goals of this theory are to serve as a simple formalism for
gaining deeper insight into the physical properties seen in detailed simulations as well as to serve
as a model to easily compare why these simulations suggest a different kinetic mechanism than
previous simple models. Specifically, we find that non-native contacts play a key role in determining
the mechanism, which can shift dramatically as the energetic strength of non-native interactions is
changed. For protein-like non-native interactions, our model finds that the native state is a kinetic
hub, connecting the strength of relevant interactions directly to the nature of folding kinetics.
PACS numbers:
Introduction.- Protein folding has been an important
problem at the crossroads of statistical mechanics, com-
puter simulation, and biophysics. There has been a long
history of theoretical advances in the study of protein
folding, and we refer the reader to reviews [1–4]. Re-
cent advances in computer simulations have enabled one
to use detailed, atomistic models to simulate the com-
plete process of folding, on relatively long (millisecond)
timescales [5]. This has become possible due to the ad-
vent of Markov State Models (MSMs) (see Refs.
[4, 6]
for recent reviews), an approach which uses detailed sim-
ulation to construct a Master equation for the statistical
dynamics of a particular protein.
By examining and comparing MSMs for different pro-
teins (as well as by direct examination of simulations),
some surprises have emerged. Perhaps most importantly,
the role of non-native contacts has now been highlighted
as a key part of protein folding [5, 7]; in hindsight, this is
natural, since amino acid interactions are not particularly
specific and there often is little free energetic difference
between say a native-like interaction between two aro-
matic residues vs a non-native like interaction [5]. This
opens a new door to re-examine simple models of pro-
tein folding and to develop a new theoretical formalism
to more naturally include non-native interactions.
In this work, we take a Master equation approach
to dynamics, much like one does computationally with
MSMs. The key question is how to model the rate ma-
trix. Previous models for protein folding kinetics [8, 9]
(derived from models of spin glass dynamics [10]) have
made very simple approximations for the nature of the
rate matrix. Here, we develop a new theoretical frame-
work for the rate matrix which allows for a more detailed
model of kinetics, especially the natural inclusion of non-
native interactions. The application of this model to var-
ious regimes of protein-sequences allows one to make a
direct connection to recent simulations [5] and, via a sim-
ple, solvable analytic model, describe the essence of na-
tive hubs (i.e. transitions between non-native states occur
via the native state) recently seen in simulation [7].
Model.- We first introduce a phenomenological Hamil-
tonian for the energy of structure α:
Hα = ǫN X
ij
Cα
ij CN
ij + ǫN N X
ij
Cα
ij(1 − CN
ij )
(1)
where Cα
ij is the contact map of structure (microstate)
α (i.e. either 1 if a contact between residues i and j is
present in the structure α or 0 otherwise) and CN
ij is the
contact map of the native state. We choose the native
contacts and non-native contacts to have differing ener-
getic contributions (ǫN and ǫN N , respectively), noting
that these quantities could naturally be negative (espe-
cially ǫN ). Note that we are including terms such as
solvent entropy in the "energy" term above.
This Hamiltonian reduces
to Hα = ǫN qα,N +
ǫN N (qα,α − qα,N ) = ǫxqα,N + ǫN N qα,α, where ǫx ≡
ǫN − ǫN N is the extra energetic preference for native over
non-native contacts and qα,β ≡ Pij Cα
ij is the number
of contacts in common between structures α and β (also
note that thus qα,α is shorthand for the total number of
contacts in structure α). This Hamiltonian is simple, yet
captures the main element of interest in this work: the
interplay between native and non-native interactions.
ij Cβ
However, in order to make a connection to more de-
tailed calculations,
it is useful to note that a simi-
lar expression can be derived from more direct physi-
from a microscopic Hamiltonian with
cal grounds, i.e.
the explicit concept of sequence design [2].
Starting
from the more general, microscopic contact Hamiltonian
Hα = Pij Cα
ij Bsi,sj , where BIJ is a general matrix of
monomer-monomer interactions (and we use capital let-
ters to designate the space of different types of residues).
This is analogous to problems that have already been
solved [2], yielding (to lowest order in 1/Td) the effective
Hamiltonian for α is
Heff
α = Fα + T Sloop
α = Bqαα −
B2
c
2Td
qαN
(2)
where Td is the design temperature (lower Td means bet-
ter optimized sequences), and we see terms involving
the mean (B) and the variance (B2
c ) of the BIJ ma-
trix take the roles of of ǫN N and ǫx, respectively,
in
our phenomenological Hamiltonian. Having both rep-
resentations (i.e.
the phenomenological as well as the
microscopic Hamiltonian) allows us also to make a natu-
ral connection to previous work [2]. We will return to the
sequence-based Hamiltonian results at the end, in order
to make a more direct connection to protein biophysics.
Kinetics formalism.- To build a kinetic model, we con-
sider the master equation:
dpα
dt
= X
β
kαβ pβ
(3)
which means that we must consider the nature of the
rate matrix kαβ. To more easily see the impact of these
elemental rates on the overall dynamics, we propose a
simple model for the rate matrix: a block-diagonal ma-
trix with a block diagonal form of n blocks each of size
m rows but now with one additional row for the folded
(native) state. Specifically, with elements of the form
2
where k is the microscopic rate of interconversion, Fα is
the free energy of α, and F ‡
αβ is the free energy of the
transition state (denoted by ‡) between α and β (note
that this is not the global transition state, but just the
transition state between structures α and β). Since the
eigenvalues and eigenvectors of the kαβ matrix define the
relevant timescales and dynamics, respectively, the next
step is to flesh out this matrix in more detail.
There have been previous approximations to model
kαβ, notably setting kαβ = k exp(−Eβ/T ) [10] or kαβ =
k exp(Eα/T ) [11], which yielding solvable models within
the Random Energy Model (REM) leading to power-
law [10] and stretched exponential [11] relaxation, re-
spectively. Also of note is an extension to GREM [8].
However, there are two key limitations to these methods.
First, they only directly apply to a theory for kinetics of
random sequences. Second, as we argue below, by consid-
ering the structure of the transition state for transitions
explictly, we can improve upon the previous models.
We propose that the transition state between struc-
tures α and β can be approximated in terms of the con-
tacts in common between these structures:
k1
k0
k0N
kN 0
within a nonnative block
between nonnative blocks
from non−native to native
from native to non−native
kαβ =
−Pβ6=α kαβ on diagonal
(4)
Here, the n blocks are meant to represent n mestastable
states, each consisting of m highly related conformations.
Note that this matrix obeys detailed balance, although
we have made the simplifying approximation that the free
energy of states within a block are similar and the free
energy between non-native blocks is also similar.
This rate matrix has a well defined set of degener-
ate eigenvalues: a non-degenerate eigenvalue of 0 (the
equilibrium eigenvalue), a (n − 1)-fold degenerate eigen-
value of κ0 ≡ nmk0 + kN 0 (for transitions between non-
native blocks), a [n(m − 1)]-fold degenerate eigenvalue of
m(n − 1)k0 + mk1 + kN 0) (for transitions within a block),
and a non-degerate eigenvalue of κN ≡ nmk0N +kN 0 (for
transitions to the native state). We note in passing that
the degeneracy in the eigenvalues seen here is naturally
broken by some small variations in rates between states
(i.e. the rates between all non-native states would not be
all exactly k0). Also, variations in the value of m between
blocks do not change the results discussed below.
In the large n limit, the ratio of rates of transforma-
tions between non-native blocks vs those from non-native
to native will be κ0/κN = (nmk0 + kN 0)/(nmk0N +
kN 0) ≈ k0/k0N . Thus, our primary goal will be to com-
pare these elemental rates. To do so, we take a Kramer's
approximation for dynamics between α and β, i.e.
kαβ = k exp[−(F ‡
αβ − Fα)/T ]
(5)
C ‡
ij = Cα
ij Cβ
ij
(6)
(here, we take advantage of the fact that contacts are
either valued at 0 or 1, so multiplication works like a
binary AND operator). Physically, this models the tran-
sition between α and β as breaking the contacts in α not
present in β and then forming the contacts present in β
that were not originally present in α. We note in pass-
ing that this approach is potentially broadly applicable
to a range of problems, whose state information can be
encapsulated into a binary vector analogous to Cα
ij .
Eq (6) is an advance over previous work in two ways.
First, we consider directly the microscopic transitions be-
tween states, i.e. not considering these transitions in
terms of all going through a single barrier, but many
different pair-wise barriers. Second, we look directly to
structural properties of the state to determine the na-
ture of the transition state structure. However, we stress
that eq (6) is most appropriate for transitions between
collapsed (or mostly collapsed) states.
To calculate the free energy as a function of a given
state, we must include the energy (from the Hamiltonian
above) as well as a model for the polymeric entropy. We
follow the model described in [12] and say that for a chain
of N persistence lengths, the number of contacts present
(qα,α) in the structure α lead to qα,α loops, each of length
ℓ ∼ N/qα,α; these loops each contribute an entropy of
/kB = −σ+(3/2) ln qα,α, where σ ≡ s−(3/2) ln N
∆Sloop
and s is a positive quantity related to the flexibility of
the chain and kB is Boltzmann's constant; note that the
value of ∆Sloop
(eg as seen in lattice model calculations
[2]) is dominated by the σ term.
α
α
This leads to the transition state energy
E‡
αβ = ǫN X
ij
ij Cβ
Cα
ij CN
ij + ǫN N X
ij
ijCβ
Cα
ij (1 − CN
ij )
= ǫxqα,β,N + ǫN N qα,β
(7)
where qα,β,N ≡ Pij Cα
ij is the three-conformation
overlap between α, β, and the native state. Combining
terms above (and including the entropy), we get
ij CN
ij Cβ
F ‡
αβ − Fα = ǫx(qα,β,N − qα,N ) + fN N (qα,β − qα,α) +
−(3/2)kBT (qα,β ln qα,β − qα,α ln qα,α) (8)
where we have defined the effective free energy of contact
formation fN N = ǫN N + kBT σ. With this barrier height
now directly connected to properties of the Hamiltonian,
we are now ready to examine specific models for folding.
Exploring the model.- To examine the model, we con-
sider some limiting cases below. First, we wish to con-
sider a model for protein-like sequences, i.e. those which
resulted from evolution (or alternatively protein sequence
design).
In order to obtain reasonable estimates for
the key transition rates in our model, i.e. k0 = kαβ
vs k0N = kαN (where α and β are representative non-
folded structures and N is the native state), we look
to equations (3), (5), and (8). First, we consider the
case ǫN < ǫN N < 0, i.e. both native and non-native
contacts are energetically preferred, but native more so.
The intrastate conversion rate will be the fastest rate,
since the states are very similar (i.e. qα,β ≈ qα,α and
qα,β,N ≈ qα,N ), which leads to a very low barrier height
from eq (8).
In order to compare k0N to k0, consider
that both ǫN and ǫN N are negative, but the drive to
form native contacts is stronger; thus, the barrier height
determined in eq (8) will be lower for transitions to N .
More specifically, it is instructive to compare the bar-
rier transitions from α to some other non-folded struc-
ture β vs folding from α to N . The difference in bar-
rier heights ∆∆F ‡ ≡ (F ‡
αN − Fα) =
∆∆E‡ − T ∆∆S‡ = −kBT ln(kαβ /kαN ) is given by the
combination of an energetic
αβ − Fα) − (F ‡
∆∆E‡ = −ǫx(qα,N − qα,β,N ) − ǫN N (qα,N − qα,β)
(9)
and entropic ∆∆S‡ = (S‡
αβ − Sα) − (S‡
αN − Sα)
− T ∆∆S‡ = −kBT σ(qα,N − qα,β) +
(10)
(3/2)kBT (qα,N ln qα,N − qα,β ln qα,β)
contributions.
3
three-body overlaps are much more rare than two-body
overlaps in the unfolded state, then qα,N > qα,β,N . Fi-
nally, due to the nature of our Hamiltonian, it would
be rare for two structures at random to have more con-
tacts in common with each other than with the native
state; thus, we can make the approximation that at least
qα,N ≈ qα,β, or more likely qα,N ≥ qα,β. Thus, putting
this all together, we find that, for this regime, ∆∆F ‡
is positive, which yields k0N > k0. As the strength of
non-native interactions gets more attractive (i.e.
ǫN N
more negative) and the difference between native and
non-native strength grows (i.e. ǫN and ǫN N negative and
ǫx < 0), the more that k0N is greater than k0, emphasiz-
ing this effect.
Moreover, since the different collapsed globules have
roughly the same number of total contacts, the rates
for unfolded state globule folding to the native state are
largely uniform (justifying our model as constructed in
Eq (4)), yielding single exponential kinetics even though
there are many parallel pathways.
Due to the exponential nature of the relationship be-
tween rates and free energies, this leads to the relation-
ship k0N > k0. What are the implications of this? In
this limit, we would find that folding to the native state
is fast, compared to dynamics from one non-folded state
to another. This makes the folded state a kinetic hub, i.e.
transitions between states are typically mediated through
the native state. Moreover, generalizations of this model
with a higher level hierarchy show (as could be seen from
numerical solutions of this model) that the native state is
a kinetic hub, i.e. transitions between non-native states
usually go through the native state.
Finally, we use the previous model to examine another
regime, where native contacts dominate, i.e. ǫN is neg-
ative but fN N is positive. In this case, native contacts
are strongly preferred and non-native contacts are dis-
couraged (eg with excluded volume repulsion and no at-
traction due to interactions, as is common in computer
simulations of G¯o models [3]). Specifically, as we see di-
rectly from eq (9), when fN N is sufficiently positive, we
obtain a negative value for ∆∆F ‡. Thus, this would lead
to the model parameters of the form k0 ≫ k0N .
Physically, interconversion between non-native states
is fast, since they are not separated by barriers (their
transition state energies have no contributions from the
free energy of breaking non-native contacts). This regime
leads to a very different picture (akin to the "smooth
energy landscape" picture previously described [3]), in
which the unfolded state interconverts quickly in general,
waiting for the rare chance to fold.
Consider the regime where both ǫx and fN N are neg-
ative quantities. Note that fN N represents the effective
free energetic drive to form contacts in general and is
negative when there is a sufficient general attraction that
beats out the loss of entropy of forming contacts. Since
Discussion and Conclusions.- While previous theoret-
ical approaches [8, 9] have made seminal contributions
to the theoretical framework for understanding folding,
these models did not model the transition state struc-
turally, which has particularly important implications for
a)
b)
3
5
6
1
4
6
4
2
N
5
3
2
1
N
FIG. 1: Two different kinetic regimes result from our theory,
as demonstrated in a simple numerical example with n = 3
blocks and m = 2 states per block, plus the native state (7
states total); see eq. (4) for details of the rate matrix struc-
ture. The theory presented here is used to calculate the mean
first passage time (MFPT) between states; edges are shown
as solid lines if the MFPT is fast (< 30/k), with the MFPT
of an edge listed to one signficant digit. In both examples,
we set kN 0 = k0N exp(−8/0.6) and k1 = 1k. a) For estimates
based on the MJ matrix regime (k0N = 0.05k, k0 = 0.001k),
we see that the native state (N) is a kinetic hub. b) For a G¯o
model regime (k0N = 0.005k, k0 = 0.5k) we see that there is
fast interconversion between unfolded states (1-6), with slow
interconversion to the native state (shown by dotted lines).
the impact of non-native interactions. As we have seen
above, the inclusion of non-native interactions critically
changes the qualitative behavior of the model; indeed,
this regime has been shown to be particularly relevant in
recent all-atom protein folding simulation.
Moreover, we can derive estimates for our parame-
ters from previous studies of proteins. For example,
the Miyazawa and Jernigan [13] matrix's mean (B ≈
c )1/2 ≈ 1.5kBT ),
−3.2kBT ) and a standard deviation ((B2
respectively. These results, consistent with other such
estimates (such as amino acid solvation free energies
[13]), combined with estimates that kTd ≈ 0.8kBT and
kBT s ≈ 1.5kBT [2] indicate that fN N < 0; our the-
ory therefore predicts that proteins would fold with the
native state as a kinetic hub (i.e. fast folding to the na-
tive state, compared to equilibration between unfolded
states) [7], depicted in Fig. 1. These averages are formally
weighted by the amino acid composition [2] of the protein
sequence and a uniform composition is used above.
We also note in passing that while an overall collapsed
model is handled by our theory, the limiting case of min-
imal (or repulsive) non-native interactions is not, since
that regime would not lead to collapsed configurations
and thus eq (6) would not be valid; however, this regime
is already well understood: the preponderance of con-
tacts are native in this case, and thus folding proceeds
by the formation of these contacts, as seen previously [2].
With this new formalism, we are able to recover and
potentially explain the behavior seen in all-atom simu-
lations [5, 7]. Specifically, we get the primary result
4
that the dynamics of interconversion from one non-folded
state to another can be very slow. This also leads to a
secondary result that native state is a kinetic hub when
there is some non-native attraction. This suggests that
simple, previous choices for a single dimensional reaction
coordinate (such as using the number of native contacts)
can lead to a misconception in terms of the fundamen-
tal dynamics of proteins, since these approaches assume
that the unfolded state is rapidly interconverting. This
is correct for some G¯o models of protein folding, but not
for models which include non-native attraction.
Finally, we stress that the property of the unfolded
state predicted from this theory does not apply to the
chemically denatured state, in which most experiments
probing the "unfolded state" of proteins have been per-
formed; our theory predicts that experiments directly ex-
amining the true unfolded state will see a much slower
relaxation time compared to the denatured state.
To conclude, one of the motivations of this work was to
develop a model which was simple enough that it could
be solved analytically, but with the key essence of pro-
tein folding seen in detailed simulations. The qualitative
change which derives from the simple addition of the role
of non-native contacts shows how this model can easily
be used to probe folding dynamics. By combining de-
tailed simulations with analytic approaches, insight in a
single system studied by simulation could be extended to
a broad range of proteins and protein folding phenomena.
Acknowledgements.- We
thank S. Bacallado, K.
Beauchamp, G. Bowman, J. Chodera, J. England, A.
Grosberg, P. Kasson, M. Levitt, L. Maibaum, and V.
Voelz for discussions and NSF (EF-0623664) and NIH
(R01-GM062868) for funding.
[1] K.A. Dill, S.B. Ozkan, M.S. Shell, T.R. Weikl. Annu.
Rev. Biophys. 37, 289-316 (2008).
[2] V. S. Pande, A. Y. Grosberg, and T. Tanaka. Rev. Mod.
Phys. 72, 259-286 (2000).
[3] J. Onuchic, Z. Luthey-Schulten, P. G. Wolynes. Ann.
Rev. of Phys. Chem. 8, 545-600 (1997).
[4] F No´e, S Fischer. Transition networks for modeling the
kinetics of conformational change in Curr. Opin. In
Struct. Bio., 18, 154-62 (2008).
[5] V. A. Voelz, G. R. Bowman, K. Beauchamp, and V. S.
Pande. J. Am. Chem. Soc. 132, 1526-1528 (2010).
[6] G. R. Bowman, K. A. Beauchamp, G. Boxer, and V. S.
Pande. J. Chem. Phys., 131, 124101 (2009).
[7] G. R. Bowman and V. S. Pande. Proc. Nat. Acad. Sci.,
USA, submitted (2010).
[8] J. Wang, J.G. Saven, P.G. Wolynes. J. Chem. Phys., 105,
11276 (1996).
[9] E.I. Shakhnovich, A.M. Gutin. Europhys. Lett., 9, 569-
574 (1989).
[10] G. J. M. Koper and H. J. Hilhorst. Europhys. Lett. 3,
1213-1217 (1987).
[11] C. De Dominicis, H. Orland and F. Laine. J. Physique
[13] S Miyazawa, and RL Jernigan. J. Mol. Bio., 256, 623-44
Lett. 46, 463-466 (1985).
(1996).
[12] V. S. Pande, A. Grosberg, and T. Tanaka. Fold. Des., 2,
109-14 (1997).
5
|
1906.07605 | 1 | 1906 | 2019-06-18T14:24:57 | Hypothesis of collective state in biomolecules and discussion on the possibility of its experimental verification | [
"physics.bio-ph"
] | The analogy between the phenomenon of live and collective states in Condensed Matter was supposed by several well-known physicists. The possibility of experimental verification of this hypothesis is discussed. | physics.bio-ph | physics | Hypothesis of collective state in biomolecules and discussion
on the possibility of its experimental verification
Anatoly M. Smolovich
Kotel'nikov Institute of Radio Engineering and Electronics (IRE) of the Russian Academy of
Sciences, Moscow, Russia
e-mail: [email protected]
Abstract: The analogy between the phenomenon of live and collective states in Condensed Matter was
supposed by several well-known physicists. The possibility of experimental verification of this hypothesis
is discussed.
In the Forward to the first edition of his monograph Superfluids [1] F. London supposed
the possibility of some quantum behavior akin to superfluidity that might play a role in
biological processes. He said that certain actions between macromolecules in biochemistry could
be understood unless they could be conceived by some quantum mechanism involving the
system as whole. It is conceivable that in some biological processes the concept of fluid state of
entropy zero could play a decisive role, for it combines the characteristic stability of quantum
states with the possibility of motion (i.e. of matter transfer and change of shape) without
necessarily implying dissipation processes. In [2] W.A. Little paid attention on this remark of
Fritz London and noted that if this should be the case, an entirely new and important
consideration would be added to the problem of understanding living systems. Then, W.A. Little
suggested that superconductivity near room temperature might be achieved in an organic
polymer, which was loosely related to the structure of DNA. In [3] H. Frohlich conjectured that
some collective quantum effects can take place in living systems. My hypothesis is that some
special type of macroscopic quantum phenomena (also called collective states) is the basic
property the life phenomenon itself. The question is how to verify this hypothesis in experiment.
This is the subject of a following discussion.
the Fermi energy. The superconducting energy gap gives stability
I suppose that here the presence of energy gap can be used as an indicator of some
collective state presence. For superconductors the energy gap is a region of suppressed density of
states around
to
superconductivity because it can be considered as a portion of energy that should be applied to
destroy superconductivity. The size of the energy gap indicates the energy gain for two electrons
upon formation of a Cooper pair. BCS theory predicts that the size Δ of the superconducting
energy gap for conventional superconductors Δ(T=0) = 1.76 kBTc, where Tc is the temperature
superconducting transition and kB is Boltzmann constant [4]. The energy gap width of high
temperature superconductors [5] and charge density wave (CWD) [6] differs but also have an
order about kBTa, where Ta is the temperature of superconducting or Peierls transition,
respectively.
My hypothesis is that the life phenomenon is also characterized by presence of some
correspondence energy gap. We don't know is it true or not but we can try to search for this
energy gap in an experimental investigation. I suppose that the width of this energy gap has an
order of kBTV, where TV is some temperature typical for living organisms, for example 300K.
This gives the estimation for energy gap about 0.026 eV.
Some simple organisms should be chosen for this research. Indeed, even simple living
organisms like unicellulate, prokaryotes or viruses are significantly more complicated than the
majority of physical research subjects. All organisms contain DNA or RNA, which store the
biological information. The simplest organisms are viruses, which consist of nucleic acid (DNA
or RNA) surrounded by a protective coat of protein called a capsid. However, opinions differ on
whether viruses are a form of life or organic structures that interact with living organisms. They
have been described as "organisms at the edge of life." In [7] viruses are considered to act an
important part in the evaluation process. The viruses don't demonstrate the life characters as
metabolism, replications, etc. when they are out of the host-cell. This allows us to use the
phage's state inside and outside the host-cell as markers of alive (V, vita) and dead (M, mort)
states [8]. Besides, viruses can be simply inactivated, for example by UV illumination or by
heating. The inactivated virus doesn't demonstrate the life characters inside the host-cell. We
don't know exactly where the border between V and M virus states is situated. We can't a priory
exclude that this border is located between the normal and inactivated states of virus. The
electronic energy levels of these states should be compared. The electronic density of states
(DOS) of the viruses inside and outside of the host-cell should be measured.
Now, let's discuss what experimental methods can be used for identification of the
energy gap. Tunnelling spectroscopy is the appropriate technique for measuring the DOS
including identification of the superconducting and CDW energy gaps [9, 10]. Scanning
tunnelling spectroscopy (STS) is very useful for various molecular objects and nanoparticles
deposited on substrates. Also, there were attempts of single DNA molecules STS investigation
across the helices [11-16]. However, its clear interpretation was inhibited by technical hurdles
including a great data spread. There is another problem. As I have mentioned, it is desirable to
compare the results of spectroscopic measurements of the phages outside and inside of the host-
cells. However, the possibility of STS investigation of phages inside the host-cell casts doubt.
The other method I would like to mention is the traditional optical spectroscopy in
different wavelength region of the electromagnetic radiation spectrum. The first spectroscopic
measurement of the superconducting energy gap was made with microwaves [17] and far
infrared techniques [18]. This spectroscopy also was used for investigations of biomolecules
[19-22]. Note, that interpretation of the biomolecules optical spectrum is very difficult task due
to the complicated structure of these molecules. The additional problem is that the size of light
spot significantly exceeds the thickness of DNA molecule. This leads to necessity of desired
signal measuring against the strong background noise.
I see the way for overcoming these difficulties in using of both STS and optical
spectroscopy methods for the same DNA molecules. Comparison of the results obtained by these
methods should help in their interpretation. The other point is comparison of the spectroscopy
results for V and M virus states. The zone of spectra where some difference between V and M
virus states will be found should be the subject of farther detailed investigations.
For decision of research methodology let's pay attention on publications [23-25] where
the virus UV inactivation was investigated. The bacteriophages Lambda (λ) and phi X 174
(ΦX174) were used in the experiments. The virus inactivation after UV irradiation was checked
by phages seeding on the corresponding strains of Escherichia coli (host-cell).
Now, I would like to propose the following research methodology. Both phages and
corresponding host-cells should be investigated to find the difference in the DOS between V and
M states of phages. The border between V and M states of phages can be situated between the
normal and inactivated phages or between the normal phage inside and outside the host-cell. We
should check both options. Firstly, the spectroscopy of host-cells should be performed before
their infecting by phages. The spectroscopy of phages in normal state should be performed both
outside and inside the host-cell. The spectroscopy of phages outside the host-cell should be done
by two different methods: STS and optical spectroscopy. In [26] the discrimination of bacteria
and bacteriophages by Raman spectroscopy and surface-enhanced Raman spectroscopy was
demonstrated. In [27] detection and discrimination of viruses was fulfilled by the use of Fourier
transform infrared spectroscopy. Also see the subsequent publications [28-33]. The spectroscopy
of phages inside the host-cell should be done only by optical spectroscopy. The same
measurement of the UV inactivated phages should be done. The DOS in V and M states of the
phages should differ even in case if the hypothesis of energy gap existence is wrong. The
difference between the DOS of V and M phages states can be found by comparison of the
measured data. The specific energy gap of living state can be possibly identified by this way.
In summary, the possibility of experimental verification of hypothesis of some
macroscopic quantum state in living organisms was discussed. I thank A.A. Sinchenko and S.V.
Chekalin for useful consulting and discussions.
REFERENCES
1. London F. Superfluids, John Wiley & Sons, Inc., New York, 1950, Vol. 1.
2. Little, W. A. Possibility of synthesizing an organic superconductor. Physical Review
134.6A (1964): A1416.
3. Fröhlich H. Long‐range coherence and energy storage in biological systems.
International Journal of Quantum Chemistry 2.5 (1968): 641-649.
4. Schmidt V. V. The physics of superconductors: Introduction to fundamentals and
applications. -- Springer Science & Business Media, 2013, Section 6.
5. Timusk T., & Statt B. (1999). The pseudogap in high-temperature superconductors: an
experimental survey. Reports on Progress in Physics, 62(1), 61.
6. Monceau P. (2012). Electronic crystals: an experimental overview. Advances in Physics,
61(4), 325-581.
7. Koonin E. V., Wolf Y. I., & Katsnelson M. I. (2017). Inevitability of the emergence and
persistence of genetic parasites caused by evolutionary instability of parasite-free states.
Biology direct, 12(1), 31.
8. Ivanitskii, G. R. (2010). 21st century: what is life from the perspective of physics?.
Physics-Uspekhi, 53(4), 327.
9. Giaever I. (1960). Energy gap in superconductors measured by electron tunneling.
Physical Review Letters, 5(4), 147.
10. Latyshev Y. I. (2008). Interlayer tunneling in stacked junctions of high temperature
superconductors, CDW materials and graphite. In Electron Transport in Nanosystems
(pp. 155-177). Springer, Dordrecht.
11. Iijima, M., Kato, T., Nakanishi, S., Watanabe, H., Kimura, K., Suzuki, K., & Maruyama,
Y. (2005). STM/STS study of electron density of states at the bases sites in the DNA
alternating copolymers. Chemistry letters, 34(8), 1084-1085.
12. Lindsay, S. M., Li, Y., Pan, J., Thundat, T., Nagahara, L. A., Oden, P., ... & White, J. W.
(1991). Studies of the electrical properties of large molecular adsorbates. Journal of
Vacuum Science & Technology B: Microelectronics and Nanometer Structures
Processing, Measurement, and Phenomena, 9(2), 1096-1101.
13. Xu, M. S., Tsukamoto, S., Ishida, S., Kitamura, M., Arakawa, Y., Endres, R. G., &
Shimoda, M. (2005). Conductance of single thiolated poly (GC)-poly (GC) DNA
molecules. Applied Physics Letters, 87(8), 083902.
14. Shapir E., Cohen H., Calzolari A., Cavazzoni C., Ryndyk D. A., Cuniberti G., ... &
Porath D. (2008). Electronic structure of single DNA molecules resolved by transverse
scanning tunnelling spectroscopy. Nature Materials, 7(1), 68.
15. Ryndyk D. A., Shapir E., Porath D., Calzolari A., Di Felice R., & Cuniberti G. (2009).
Scanning tunneling spectroscopy of single DNA molecules. Acs Nano, 3(7), 1651-1656.
16. Shapir E., Sagiv L., Molotsky T., Kotlyar A. B., Felice R. D., & Porath D. (2010).
Electronic structure of g4-DNA by scanning tunneling spectroscopy. The Journal of
Physical Chemistry C, 114(50), 22079-22084.
17. Biondi M. A., Garfunkel M. P., & McCoubrey A. O. (1957). Microwave measurements
of the energy gap in superconducting aluminum. Physical Review, 108(2), 495.
18. Glover R.E. and Tinkham M., Conductivity of Superconducting Films for Photon
Energies between 0.3 and 40kTC. Phys. Rev. 108, 243 (1957). Phys. Rev., 108, 243.
19. Letokhov V.S. Section 1 in Laser picosecond spectroscopy and photochemistry of
biomolecules: Edited by V.S. Letokhov, Adam Hilger, Bristol, 1987.
20. Telser J., Cruickshank K. A., Morrison L. E., Netzel T. L., & Chan C. K. (1989). DNA
duplexes covalently labeled at two sites: synthesis and characterization by steady-state
and time-resolved optical spectroscopies. Journal of the American Chemical Society,
111(18), 7226-7232.
21. Elhadj S., Singh G., & Saraf R. F. (2004). Optical properties of an immobilized DNA
monolayer from 255 to 700 nm. Langmuir, 20(13), 5539-5543.
22. Downes, A., Mouras, R., & Elfick, A. (2010). Optical spectroscopy for noninvasive
monitoring of stem cell differentiation. BioMed Research International, 2010.
23. Gurzadyan G. G., Nikogosyan D. N., Kryukov P. G., Letokhov V. S., Balmukhanov T.
S., Belogurov A. A., & Zavilgelskij G. B. (1981). Mechanism of high power picosecond
laser UV inactivation of viruses and bacterial plasmids. Photochemistry and
photobiology, 33(6), 835-838.
24. Nikogosyan D. N., Oraevsky A.A. and Zavilgelsky G.B. (1986) Picosecond laser UV
inactivation of bacteriophage λ and various Escherichia coli strains. Photobiochem.
Phofobiophys. 10, 189-198.
25. Nikogosyan D. N. and Zavilgelsky G.B. Section 8 in Laser picosecond spectroscopy and
photochemistry of biomolecules: Edited by V.S. Letokhov, Adam Hilger, Bristol, 1987.
26. Goeller, L. J., & Riley, M. R. (2007). Discrimination of bacteria and bacteriophages by
Raman spectroscopy and surface-enhanced Raman spectroscopy. Applied spectroscopy,
61(7), 679-685.
27. Vargas, C. A., Wilhelm, A. A., Williams, J., Lucas, P., Reynolds, K. A., & Riley, M. R.
(2009). Integrated capture and spectroscopic detection of viruses. Appl. Environ.
Microbiol., 75(20), 6431-6440.
28. Hamasha, K., Mohaidat, Q. I., Putnam, R. A., Woodman, R. C., Palchaudhuri, S., &
Rehse, S. J. (2013). Sensitive and specific discrimination of pathogenic and
nonpathogenic Escherichia coli using Raman spectroscopy -- a comparison of two
multivariate analysis techniques. Biomedical optics express, 4(4), 481-489.
29. Vishnupriya, S., Chaudhari, K., Jagannathan, R., & Pradeep, T. (2013). Single‐Cell
Investigations of Silver Nanoparticle -- Bacteria Interactions. Particle & Particle Systems
Characterization, 30(12), 1056-1062.
30. Chen, J., Zhang, Y., & Kohler, B. (2014). Excited states in DNA strands investigated by
ultrafast laser spectroscopy. In Photoinduced phenomena in nucleic acids II (pp. 39-87).
Springer, Cham.
31. Tayyarcan, E. K., Soykut, E. A., & Boyaci, I. H. (2018). A Raman-spectroscopy-based
approach for detection and discrimination of Streptococcus thermophilus and
Lactobacillus bulgaricus phages at low titer in raw milk. Folia microbiologica, 63(5),
627-636.
32. de La Harpe, K., Kohl, F. R., Zhang, Y., & Kohler, B. (2018). Excited-State Dynamics of
a DNA Duplex in a Deep Eutectic Solvent Probed by Femtosecond Time-Resolved IR
Spectroscopy. The Journal of Physical Chemistry A, 122(9), 2437-2444.
33. Li, R., Dhankhar, D., Chen, J., Krishnamoorthi, A., Cesario, T. C., & Rentzepis, P. M.
(2019). Identification of Live and Dead Bacteria: A Raman Spectroscopic Study. IEEE
Access, 7, 23549-23559.
|
1506.08051 | 1 | 1506 | 2015-06-26T13:12:33 | Mechanical properties of branched actin filaments | [
"physics.bio-ph",
"q-bio.SC"
] | Cells moving on a two dimensional substrate generate motion by polymerizing actin filament networks inside a flat membrane protrusion. New filaments are generated by branching off existing ones, giving rise to branched network structures. We investigate the force-extension relation of branched filaments, grafted on an elastic structure at one end and pushing with the free ends against the leading edge cell membrane. Single filaments are modeled as worm-like chains, whose thermal bending fluctuations are restricted by the leading edge cell membrane, resulting in an effective force. Branching can increase the stiffness considerably; however the effect depends on branch point position and filament orientation, being most pronounced for intermediate tilt angles and intermediate branch point positions. We describe filament networks without cross-linkers to focus on the effect of branching. We use randomly positioned branch points, as generated in the process of treadmilling, and orientation distributions as measured in lamellipodia. These networks reproduce both the weak and strong force response of lamellipodia as measured in force-velocity experiments. We compare properties of branched and unbranched networks. The ratio of the network average of the force per branched filament to the average force per unbranched filament depends on the orientation distribution of the filaments. The ratio exhibits compression dependence and may go up to about 4.5 in networks with a narrow orientation distribution. With orientation distributions measured in lamellipodia, it is about two and essentially independent from network compression, graft elasticity and filament persistence length. | physics.bio-ph | physics |
Mechanical properties of branched actin filaments
Mohammadhosein Razbin1, Martin Falcke2,
Panayotis Benetatos3 and Annette Zippelius1
1 Max Planck Institute for Dynamics & Selforganization, Am Fassberg 17 and
Institute for Theoretical Physics, Georg August University, Friedrich-Hund-Platz 1,
37077 Gottingen, Germany;
2Max Delbruck Center for Molecular Medicine, Robert Rossle Str. 10, 13092 Berlin,
and Dept. of Physics, Humboldt University, Newtonstr. 15, 12489 Berlin, Germany;
3 Department of Physics, Kyungpook National University, 80 Daehakro, Bukgu,
Daegu 702-701, Korea
E-mail: [email protected]
[email protected]
Abstract. Cells moving on a 2dimensional
substrate generate motion by
polymerizing actin filament networks inside a flat membrane protrusion. New filaments
are generated by branching off existing ones, giving rise to branched network structures.
We investigate the force-extension relation of branched filaments, grafted on an elastic
structure at one end and pushing with the free ends against the leading edge cell
membrane. Single filaments are modeled as worm-like chains, whose thermal bending
fluctuations are restricted by the leading edge cell membrane, resulting in an effective
force. Branching can increase the stiffness considerably; however the effect depends on
branch point position and filament orientation, being most pronounced for intermediate
tilt angles and intermediate branch point positions. We describe filament networks
without cross-linkers to focus on the effect of branching. We use randomly positioned
branch points, as generated in the process of treadmilling, and orientation distributions
as measured in lamellipodia. These networks reproduce both the weak and strong
force response of lamellipodia as measured in force-velocity experiments. We compare
properties of branched and unbranched networks. The ratio of the network average of
the force per branched filament to the average force per unbranched filament depends
on the orientation distribution of the filaments. The ratio exhibits compression
dependence and may go up to about 4.5 in networks with a narrow orientation
distribution. With orientation distributions measured in lamellipodia, it is about 2
and essentially independent from network compression, graft elasticity and filament
persistence length.
PACS numbers: 00.00, 20.00, 42.10
Keywords: Branched actin networks, lamellipodium, force-extension relation
Accepted in Physical Biology
Mechanical properties of branched actin filaments
2
1. Introduction
The crawling of many different cell types is essential for life. Undifferentiated cells move
towards the site, where they form a tissue or organ in the developing embryo. Skin
cells start crawling when they have to close a wound [1]. During metastasis, cancer cells
dissociate from the primary tumor, crawl towards blood vessels and spread all over the
body [2, 3]. Branched actin filaments carry forces during cell motion, and consequently
understanding their elastic properties is central to understanding the mechanics of cell
motility.
In vitro, cells are plated on a two dimensional substrate to observe their dynamics.
They form a flat membrane protrusion in the direction of motion, the lamellipodium,
which is only about 100-200 nm thick but several µm deep and wide [4]. A dense
network of actin filaments (F-actin) inside the lamellipodium pushes the leading edge
membrane forward [5]. Treadmilling of the filaments drives motion [6]: Filament barbed
(or plus) ends polymerize at the leading edge of the lamellipodium and the pointed (or
minus) ends depolymerize at the rear.
Usually cells move in response to an external signal. A variety of signals stimulate
the activation of nucleation promoting factors (NPFs) (like WASp or WAVE) located in
the leading edge membrane of the lamellipodium. They activate the actin related protein
complex Arp2/3. It binds to an existing filament very close to or at its barbed end at
the lamellipodium's leading edge. That initiates the growth of a new filament branch
out of the Arp2/3 complex. Many of these branched structures consisting of mother
filament and branch form the F-actin network in the lamellipodium. The branched
structure itself is dynamic. The branch point with the Arp2/3 complex moves rearward
due to treadmilling in the same degree as mother filament and branch grow. Since
Arp2/3 binding to the individual filaments is not synchronous we find at any time many
different positions of branch points in the lamellipodium F-actin network.
The elastic properties of the F-actin network crucially depend on the density of links
between filaments [7, 8, 9, 10, 11, 12, 13]. Molecular links arise in two ways: Cross-
links connecting two filaments at some point along their contour length are formed by
cross-linker molecules like filamin or α-actinin, and branching attaches the minus end
of a filament laterally to a mother filament.
Intuition suggests that branching alone
could stiffen the network to some degree, since branching is a geometrical constraint
on the configuration of two filaments. That intuition has never been quantified before,
but is supported by our results presented below. On the other hand, the network
region close to the leading edge was found to be as soft as weakly cross-linked actin
networks [9, 10, 11, 12, 13], and experiments in actin solutions suggest that branching
contributes very little to the elastic modulus of F-actin networks [14]. Here, we would
like to present a first step in quantifying the contribution of branching to the elastic
and semi-flexible properties of the lamellipodial F-actin network. How much stiffer than
single filaments are branched filaments? How are their properties reflected in network
behavior? We will answer these questions by investigating a single branched filament and
Mechanical properties of branched actin filaments
3
networks of branched structures in an approximation neglecting interactions between
them in order to focus on branching effects. This neglect of interactions implies that we
consider only elastic properties on short time scale and not the visco-elastic properties
arising from cross-linking.
The mother filament is grafted at one end and has a free tip at the other one
in our model system (see Fig. 1). The graft is provided by a highly cross-linked part
of the F-actin network. This idea is based on the increasing degree of cross-linking
and filament bundling towards the rear of the lamellipodium, which has been observed
in many different experiments and simulations [5, 15, 16, 17, 18, 19, 20, 21, 22, 23]
(see [24] for a detailed discussion). The graft moves in the direction of cell motion due
to cross-linker binding and bundling, and thus contour length 'flows' into the graft. In
the steadily moving cell, the balance between polymerization and cross-linking creates
a stationary distance between graft and leading edge membrane. At any time, the
branched structure is in a configuration similar to Fig. 1, but the branch point moves
with treadmilling towards the graft point. Both mother filament and branch polymerize
such that their barbed ends stay at the leading edge membrane. Consequently, we will
vary the position of the branch point on the mother filament from 0 to its full length
L when we investigate the elastic properties. We model the filaments in the weakly
bending regime, i.e. bending does not affect the end-to-end distance. We allow for
an elastic graft of stiffness Ks and model the membrane by a constraint, enforcing the
filaments to be entirely on the left side of the membrane.
After the introduction of the model, we will discuss results for a single branched
filament. The F-actin network in the lamellipodium comprises filaments with many
different tilt angles and branch point positions. Hence, we consider it as an ensemble of
branched structures with varying branch point and tilt angle and calculate its properties
as the average over this ensemble. The network average of the force will be compared
to experimental results from force-velocity measurements and the force of unbranched
networks.
2. Results
2.1. The model: semiflexible branched filaments in the weakly bending limit
We model the interior of the lamellipodium as a two-dimensional space, since it is
approximately flat as described above. Furthermore, the results are easily generalised
to three dimensions. We always assume a sufficiently large persistence length, such that
the weakly bending approximation applies and the fluctuations perpendicular to the
mean orientation of the polymer segment are small and can be treated on a Gaussian
level.
Our elemental structure is a grafted filament of contour length L. The probability
to find its tip at position (x, y) with orientation θ, given that it is grafted at (x0, y0) with
orientation ω is denoted by GL(x, y, θ(x0, y0, ω). For the simple case of perpendicular
Mechanical properties of branched actin filaments
4
Figure 1. Left: Single branched filament grafted on a soft graft and confined in
the x-direction by a flat membrane. Thermal bending fluctuations are not shown in
this drawing. L1 is the contour length between graft point and branch point, L3 is
the contour length between branch point and filament tip. The contour length of the
mother filament is L = L1 + L3, Lb is the branch contour length. The branch angle is
γ = 70◦ throughout the study, and ω is the tilt angle. The numbers 1, 2 and 3 refer
to the filament tip coordinates in Eq. 7. Right: Network of branched filaments with
various orientations and locations of branch points. The length δ0 is the distance of
the filament tips from the graft plane without bending and fluctuations. The lengths
cos(ω) , Lb = L3 cos(ω)
L and Lb obey L = δ0
cos(ω−γ) , and δ denotes the distance between the graft
point and the leading edge membrane.
grafting, ω = 0, at (x0, y0) = (0, 0), GL obeys [25]
)]δ(x − L).
The general case is obtained by a translation and rotation according to
GL(x, y, θ0, 0, 0) ∝ exp[−
3lp
L3 (y2 − Lyθ +
L2θ2
3
θ → θ − ω
y → (y − y0) cos ω − (x − x0) sin ω
x → (x − x0) cos ω + (y − y0) sin ω
and explicitly given by Eq. 16.
(1)
(2)
In a first step we compute the probability, Pt(xx0), that the endpoint of the tilted
polymer is at a distance of x − x0 from the graft point:
Pt(xx0) ≡ Z Z dydθ GL(x, y, θx0, 0, ω).
(3)
We consider an elastic, fluctuating structure into which the filament is grafted and hence
model it by a fluctuating spring in x-direction with spring constant Ks = (kBT )K,
zero equilibrium length and a distribution of the spring extension proportional to
exp(−(Kx2
0)/2). The stiffness of the substrate, K, is assumed to be large as compared
to the stiffness for transverse fluctuations, K ≫ K⊥ = 2lp
3L3 , of a filament of contour
3(L cos(ω))3 . The probability Pt(x) to find the
length L. In this study we take K = 100
filament tip at x then follows by averaging over all graft point positions x0:
2lp
Pt(x) = Z
dx0√2πK
Pt(xx0) exp(−
Kx2
0
2
).
(4)
Mechanical properties of branched actin filaments
5
The filament exerts a force on an impenetrable flat membrane at a distance δ from
the graft plane (see Fig. 1). The force originates from the reduction in the number of
filament configurations due to the constraint x − x0 ≤ δ imposed by the membrane; we
therefore call it an entropic force. The fraction of configurations satisfying the constraint
is given by
Zt(δ) = Z dxPt(x)Θ(δ − x).
The derivative of Zt with respect to δ is the entropic force [26]
ft(L, δ, ω) = kBT
∂
∂δ
ln Zt(δ).
(5)
(6)
These results for a single grafted polymer are easily generalised to the branched
structure shown in Fig. 1. The probability to find the two endpoints at x2 and x3
respectively given that the structure is grafted at point x0 is calculated by the following
expression:
Pb(x2, x3 x0) =
Z Z Z dx1dy1dθ1GL1(x1, y1, θ1 x0, 0, ω)
×Z Z dy2dθ2GLb(x2, y2, θ2 x1, y1, θ1 − γ)
×Z Z dy3dθ3GL3(x3, y3, θ3 x1, y1, θ1)
(7)
The stiffness of the graft is again taken into account by averaging over the positions of
the graft:
Pb(x2, x3) ∝ Z dx0Pb(x2, x3x0) exp(−
Kx2
0
2
)
(8)
The explicit expression for Pb(x2, x3) is given by Eq. 22 in the Appendix. The fraction
of configurations satisfying the constraint obeys in analogy to Eq. 5
Zb(δ) = Z dx2Z dx3Pb(x2, x3)Θ(δ − x2)Θ(δ − x3),
and the entropic force follows from
fb(L, Lb, L3, δ, ω) = kBT
∂
∂δ
ln Zb(L, Lb, L3, δ, ω).
(9)
(10)
The force fb depends on the parameters L ≡ (L1 + L3), L3, ω and δ (Lb is fixed by ω,
L and L3). We will present our results as dimensionless quantities, and scale lengths
by δ0 and force by kBT /δ0 for that purpose. We comment on which specific values
of parameters are suggested by experimental observations in the Discussion and the
Appendix below.
Mechanical properties of branched actin filaments
6
Figure 2. Left panel:
graft point for the symmetric case ω = 35◦ and L3 = Lb = L
versus tilt angle ω for δ = δ0. The curves represent the ratio of lp
from top to bottom in both figures.
force versus relative distance, δ/δ0, between membrane and
force
= 10, 5 and 10
3
2 . Right panel:
δ0
2.2. Properties of single branched filaments
The entropic force exerted by the branched filament on the membrane is shown in the
left panel of Fig. 2 for the symmetric case ω ∼ 35◦ and branching at the midpoint of
the mother filament. The force decreases with increasing distance δ between grafting
plane and membrane. An infinitely stiff filament (lp very large) would just touch the
membrane, if δ equals δ0 = L cos ω. But the force exerted by a semiflexible filament
is nonzero even for δ > δ0, because the tilted branched structure exhibits fluctuations
with the endpoint reaching beyond δ0. In this regime the force is nearly independent of
persistence length, whereas for δ < δ0 we observe a strong increase with lp.
The force exerted by the filament on the membrane is crucially affected by the tilt
angle. It is plotted as a function of tilt angle ω for fixed δ = δ0 in the right panel of
Fig. 2. Remarkably, there is a shoulder in all three curves around ω ∼ 35◦, indicating
that the symmetric structure generates comparatively large forces. For ω → 0, the
force would diverge for a stiff graft, since Eq. 1 excludes longitudinal fluctuations of
the filament tip. The high but finite stiffness of the graft limits the force which can
be exerted on the membrane. These two opposite effects generate the maximum in the
force close to ω = 0.
Figure 3 shows the entropic force of a branched structure for the whole range of
branch point positions on the mother filament. At the time of Arp2/3 binding to the
mother filament (Lb=0), the complete structure has of course the properties of the
mother filament. When the branch point has reached the graft plane, the branched
structure corresponds to two filaments with the corresponding tilt angles. In between,
the force has a maximum at Lb = 0.484L. The maximum force is about 2.25 times the
force of two filaments with length L. For the special case under consideration, δ = δ0,
the maximum of the force scales like the square root of lp, as shown in the right panel
of Fig. 3. For the compressed case, δ < δ0, we observe a crossover to linear scaling. To
understand this behaviour of the force, we consider the case of a single filament: The
Mechanical properties of branched actin filaments
7
force versus relative length of the daughter branch. Below: the
Figure 3. Left:
configurations of the branched filament corresponding to Lb/L = 0, 0.5, 1.0. The
parameter values are: L3 = Lb, ω = 35◦, δ = δ0. From high to low force values
the symbols represent the ratio of lp
3 , respectively. Right: log-log plot
δ0
of the maximum force as a function of persistence length lp, fitted to a square root
= 10, 5 and 10
dependence fmax ∝ plp.
force has a thermal (entropic) and a nonthermal (energetic) contribution. The thermal
force dominates for δ > δ0. Evaluating the force given in Eq 21 for δ = δ0 and large K,
b
we explicitly see f ∝ plp. As we compress the filament the energetic force becomes the
dominant contribution. It can be computed as the force to bend a grafted cantilever
beam resulting in f = ( 3lpkBT
. Hence we observe a crossover from the scaling
of the force with plp to linear scaling as δ is decreased below δ0
Both, the maximum force, fmax, as well as its branch point position, Lmax
2L3 ) (δ−L cos(ω))
sin2(ω)
, depend
on the tilt angle ω. The general dependence on ω is shown in the top left panel
of Fig. 4 for fmax and in the top right panel for Lmax
. Since the dependencies are
nonmonotonic, we plot in the bottom panel the configurations which give rise to the
maximum force and help to understand the non-monotonic dependence. For very small
ω, i.e. almost perpendicular grafting, Lmax
is very small and hence also the distance
between the branch point and the endpoint of the mother filament, denoted by Lmax
.
As the tilt angle, ω, increases, the branch point moves further away from the leading
edge membrane and hence both, Lmax
increase. As ω approaches 70◦, the
branch is almost perpendicular to the membrane and the maximum force is obtained
for branching at the grafting plane, implying Lmax
/L as in the top
right panel of Fig. 4, one actually observes a decrease of Lmax
/L, because L grows faster
than Lmax
Lmax
as ω → 70◦.
is independent of the persistence length, but does depend on δ. For
3 = L. Plotting Lmax
and Lmax
3
b
3
b
b
b
b
b
3
Mechanical properties of branched actin filaments
8
Figure 4. Upper left: Maximum of the force fmax obtained by varying Lb, versus
tilt angle ω; we have δ = δ0 for the curves. From high to low force values the curves
represent the ratio of lp
3 , respectively. Upper right: Relative branch
δ0
length Lmax
/L of the maximum force versus tilt angle ω for the same parameters
(independent of the persistence length lp). Bottom: Relative partial contour length of
the mother filament in between the endpoint and the branch point with maximal force
Lmax
cos(ω − γ)/ cos ω; also shown are three representative configurations.
= 10, 5 and 10
3 = Lmax
b
b
3
3
δ ≥ δ0 entropic contributions dominate and the force is largest for two independently
fluctuating filaments, such that Lmax
/L quickly approaches 1 as δ extends beyond δ0.
On the other hand, for δ < δ0, elastic contributions are important. In the symmetric
case, we find 0.435 ≤ Lmax
/L ≤ 0.484 for 0.5 ≤ δ/δ0 ≤ 1.0.
The maximum force is observed for ω ∼ 0 and ω ∼ 70◦, because either the mother
filament or the daughter filament are perpendicular to the membrane (upper left panel of
) . When the branch is perpendicular, the force is mainly determined by the fluctuations
of the branch point, and when the mother filament is perpendicular by the fluctuations
of the graft point. Moving away from perpendicular incidence the force has to decrease,
giving rise to a minimum for intermediate ω.
Mechanical properties of branched actin filaments
9
800
600
400
200
>
T
B
k
/
0
d
b
f
<
0
0.5
0.7
0.9
d/d
0
1.1
1.3
6
0
36
30
24
18
12
lp
/d
0
1500
1000
f
k
500
0
0.5
0.6
0.7
0.8
0.9
d/d
0
1
1.1
1.2
1.3
0
6
36
30
24
18
12
/d
lp
0
Figure 5. Left panel: Average force per branched structure of the network as a
function of δ/δ0 and lp/δ0. Right panel: The force constant kf , which is the derivative
of fbδ0/kBT with respect to δ/δ0.
R
d/d
0
Figure 6. Ratio R of the average force per filament of branched to unbranched
networks versus relative distance δ/δ0 between the membrane and the graft point;
lp
= 10. The squares refer to the random network with a uniform orientation
δ0
distribution in the range −70◦ ≤ ω ≤ 70◦ as measured in ref. [27], the circles show
results with a narrow uniform distribution of orientations in 30◦ ≤ ω ≤ 40◦, the dots
refer to the orientation distribution given by Eq. 13 as measured in ref. [28].
2.3. Extension to the F-actin network: Properties of an ensemble of branched filaments
We would like to obtain an estimate of how branching affects the network properties.
With the theoretical means set up above and in the spirit of the study, we calculate the
force as an average across an ensemble of branched structures, in which all interactions
between the branched filaments are neglected. The ensemble is described by branch
point (L3)and orientation (ω) distributions. The branch point is equally likely anywhere
on the mother filament, corresponding to 0 ≤ L3 ≤ δ0/ cos ω. The length of the branch
obeys Lb = L3 cos ω
cos(ω−γ) . As far as the orientation of the mother filaments is concerned,
several scenarios have been discussed in the literature. In ref. [27], electron microscopy
was used to determine the orientation of filaments in lamellipodial actin networks. The
distribution was found to be approximately uniform in the range of angles between 0
< fb(δ) >= Z ωmax
×Z
dL3 fb(
ωmin
δ0
cos ω
0
δ0
cos ω
,
L3 cos ω
cos(ω − γ)
, L3, δ, ω).
Mechanical properties of branched actin filaments
10
and 60◦ with a small contribution between between 60◦ and 75◦. We describe it here as
a uniform distribution between -70◦ and 70◦.
The force for fixed δ0 averaged over branch point positions and tilt angles is given
by
dω
cos ω
δ0∆ω
(11)
where ∆ω = ωmax − ωmin denotes the range of the distribution. The average force is a
monotonic function of both, the persistence length and the distance δ between grafting
plane and membrane, as shown in Fig. 5.
We can compare our results with measurements of the force-velocity relation of fish
keratocyte lamellipodia. The forces exerted by the leading edge of the freely running
cell immediately upon collision with the cantilever were below the force resolution of
the cantilevers [13, 11, 12], but caused an immediate decrease of leading edge velocity
by 1-3 orders of magnitude. Hence, the leading edge is much softer than the softest
cantilever used in the experiments, which had a force constant of 9.1 nN/µm. If the
leading edge had the same force constant, a single branched structure would need to
have a constant of about 11.8 pN/µm with 220 filaments/µm [29] and a contact length
of about 7 µm [12]. The value of δ0 in the freely running cell was estimated to be
∼1.3 µm [13], and such a value is also supported by the data in ref. [30]. Hence, if
the leading edge had the same force constant as the softest cantilever, the dimensionless
0(kBT)−1=4748
force constant of a single branched structure would be 11.8 pNµm−1δ2
(kBT=4.2 10−3 pNµm). The values in the right panel of Fig. 5 show that the branched
filaments are much softer on average, which is in agreement with the dramatic velocity
drop of the lamellipodium leading edge upon collision with the cantilever.
We do not know measurements of the pressure exerted by the filaments on the
leading edge membrane in the freely running fish keratocyte but can estimate it from
our results. Assuming δ0=1.3 µm, lp/δ0 ≈ 10, δ/δ0 ≈ 0.8 [13] we obtain f δ0/kBT ≈ 20
and a force per branched structure of 0.063 pN. The pressure exerted by the filaments
on the leading edge membrane is in the range of 46 Pa (with 110 branched structures
per micrometer and a lamellipodium height of 150 nm as in [12]).
The value of δ0 decreases during the force-velocity measurement due to a dynamic
equilibrium between polymerization and cross-linking [13].
In the stalled state,
δ0 ≈0.27 µm applies (Fig. S3 of [13]). The measured stall pressure exerted on the
leading edge by the cantilevers is 300-750 Pa [13] and 110-430 Pa [12]. These pressures
correspond to stall forces of 0.045-0.1125 nN [13] and 0.0165-0.0645 nN [12] per
micrometer leading edge. The values of f δ0/kBT resulting from these force densities
are in the range 10-70. That entails δ/δ0 . 1 (see Fig. 5) in agreement with modelling
results in ref. [13].
To assess the effects of branching on network properties, we calculate the ratio R
Mechanical properties of branched actin filaments
11
of the force exerted by a branched network to the force of an unbranched network with
the same angular distribution. We use the same number of filaments in the unbranched
network as there are branched structures in the branched network:
< fb >
dω ft(δ0/ cos ω, δ, ω)
.
(12)
R(δ) =
1
∆ω R ωmax
ωmin
The ensemble averages of branched and unbranched networks with the uniform
orientation distribution between -70◦ and 70◦ behave very similar under compression
(Fig. 6, squares). This ratio has the remarkable property of depending only very weakly
on δ. It is almost independent of the persistence length lp and the graft stiffness K
as well (data not shown). Hence, the behavior of both networks scales very similar
in dependence on these parameters. The ratio is about 2, i.e., the average force per
filament tip for a given value of δ is the same for branched and unbranched networks.
However, there are obvious qualitative differences between single filaments and
branched structures illustrated by the non-trivial dependency of the properties of
branched structures on the branch point position in Figs. 3 and 4.
if we
use a narrow orientation distribution around ω = 35◦ the value of R increases
with increasing compression. Therefore, we also investigate non-uniform measured
lamellipodial orientation distributions to investigate whether the dependency of R on
the distribution affects the behavior of lamellipodial networks. Distributions peaked
either at ω = 0 or ω = ±35◦ have been observed in refs. [28, 31, 32]. The distribution
in ref. [28], their Fig. 4, is typical for the measured non-uniform distributions and can
be approximated by
Indeed,
P (ω) = 0.008012(cid:20)e
−(ω−35)2
2·20.52 + e
+ 0.0006907, −85◦ ≤ ω ≤ 85◦.
−(ω+35)2
2·20.52 (cid:21)
(13)
The dots in Fig. 6 show the results for R. They are very similar to the results for
the uniform distribution. Obviously, the width of 40◦-45◦ of the peaks in lamellipodial
orientation distributions is too large for an essential effect of branching on the parameter
dependencies.
3. Discussion
We investigated the properties of branched filaments grafted into an elastic graft. Their
stiffness has a maximum in its dependence on the branch point position. Branched
structures with the optimal graft point position can be more than four times as stiff as
a single filament with the same tilt angle as the mother filament (Fig. 3), while requiring
only 1.5 times the polymer length of the mother filament.
The mechanical properties of branched F-actin networks depend on their orientation
distribution. With measured lamellipodial distributions, networks of branched
structures are about twice as stiff as unbranched networks. The total network force
of branched networks scales essentially the same as the one of unbranched networks
Mechanical properties of branched actin filaments
12
with the parameters F-actin persistence length, graft stiffness and compression (δ).
An intuitive explanation would be, that in the end it is the single filament behaviour
determining the stiffness for both unbranched and branched filaments, since the branch
leans on the (single) mother filament when experiencing a force.
Our theory considers individual branched filaments and thus establishes the
constitutive relations on which complex network studies including cross-linking can
be based. We considered F-actin networks as defined by their geometrical property
distributions without interactions of filaments by cross-linking or entanglement. This
implies that we cannot account for visco-elastic properties. A variety of evidences
suggests the existence of a region close to the leading edge, where cross-linking is
not dominating the network properties and to which our theory directly applies.
Measurements of the ratio of number of the cross-linkers to the number of actin molecules
in fibroblasts show the existence of a gradient for α-actinin and ABP-280/filamin.
The number ratios are low in a region juxtaposed to the leading edge with a depth
of about 1.5 µm (see Fig. 5 of ref. [15]). Svitkina and Borisy conclude from these
results and structural information from electron micrographs that the impact of ABP-
280 and α-actinin on filament cross-linking is likely to be expressed more deeply in the
cytoplasm [15]. Their statement is supported by the immediate response of the leading
edge of fish keratocytes to small forces indicating also weak cross-linking in the network
region close to the leading edge [13, 11, 12]. This suggests that while understanding
of the visco-elastic properties of the network in the lamellipodium bulk requires taking
cross-linking into account, our ensemble average is applicable to a network region close
to the leading edge. The reproduction of both the weak and strong force responses
of the lamellipodium leading edge measured in force-velocity relations by our network
calculations strongly supports that conclusion (Fig. 5).
We did not take contributions from entanglement or excluded volume effects into
account when calculating the network forces. This implies that R provides only a
meaningful approximation,
if these effects are similar in branched and unbranched
networks. To the best of our knowledge, that has not been investigated quantitatively
yet. We can only provide heuristic considerations in favour of our assumption based
on comparing a variety of simulations with and without excluded volume effects with
force-velocity measurements.
Model networks of semi-flexible filaments not taking into account excluded
volume effects reproduce the elastic properties measured in force-velocity experiments
quantitatively [13]. Schreiber et al. simulated the force-velocity relation of motile cells
with rigid rods as model filaments taking excluded volume effects into account [33]. They
found excluded volume effects to be stronger in branched than in unbranched systems.
The model network of Schreiber et al. shows a response to external forces in the force-
velocity relation at about 8 nN/µm [33]. However, the lamellipodium leading edge
exhibits elastic responses to forces smaller than 0.05 nN/µm in experiments [13, 11, 12].
Additionally, bending of filaments has been observed in lamellipodia [27, 34, 35], i.e.,
filaments do not behave like stiff rods. Hence, the lamellipodium network is likely to be
Mechanical properties of branched actin filaments
13
in a parameter regime where excluded volume effects are less relevant than suggested
by a network of stiff rods.
Branching has also been observed with microtubule [36]. The branching angle
varies between 0 ◦ and 90 ◦, and it is not known, how rigid the connection to the mother
filament is. If the branch is rigidly connected, our results should apply also to these
branched structures with the adapted persistence length (a few millimeters [37]) and
γ-values.
In summary, single branched filaments with intermediate branch point positions and
tilt angles exert larger forces and are stiffer than two unbranched filaments. The stiffness
of a whole network of branched filaments is largest, if the orientation distribution is
sharply peaked around ±γ/2 = ±35◦. For lamellipodial orientation distributions, the
stiffness of branched and unbranched networks scales approximately the same with a
variety of parameters (Fig. 6), suggesting that the effect of branching on network stiffness
can be accounted for by rescaling the filament number of an unbranched network. These
results are in agreement with the elastic properties of lamellipodia found in force-velocity
measurements with fish keratocytes [13, 11, 12].
4. APPENDIX
4.1. Tilted filament
The Green function of a free semiflexible polymer in two spatial dimensions (see Fig. 1)
and in the weakly bending limit satisfies the partial differential equation (PDE)
[
∂
∂s
+ θ
∂
∂y −
1
lp
∂2
∂θ2 ]G(s, y, θ 0, y0, ω) = 0.
The arc length of the filament contour is denoted s here. The boundary condition
lim
s→0
G(s, ys, θs0, y0, ω) = δ(θ − ω)δ(ys − y0)
(14)
(15)
realizes the graft. The solution is Eq. 1, after a switch from the coordinates (s, y, θ) to
(x, y, θ). We obtain the Green function for a tilted filament simply by a rotation:
GL(x, y, θx0, y0, ω) ∝
exp[−
lp
−
L
× exp[+
× δ[(x − x0) cos(ω) + (y − y0) sin(ω) − L].
3lp
L3 ((y − y0) cos(ω) − (x − x0) sin(ω))2
(θ − ω)2]
3lp
L2 ((y − y0) cos(ω) − (x − x0) sin(ω))(θ − ω)]
(16)
The probability distribution of the position x of the endpoint follows by integration (see
Eq. 4)
Pt(xx0) ∝ exp(−
(x − x0 − L cos(ω))2
σ2
t0
)
(17)
Mechanical properties of branched actin filaments
where
σ2
t0 =
4L3 sin(ω)2
3lp
.
14
(18)
Representing the stiffness of the graft by a spring and averaging over all grafting points,
x0, yields:
(19)
(20)
(21)
(22)
(23)
(24)
Pt(x) =
1
Nt
exp(−
(x − L cos(ω))2
σ2
t
).
Here Nu accounts for the proper normalization and
t = σ2
σ2
t0 +
2
K
.
The entropic force on the wall is then explicitly given by:
ft(δ) =
2kBT
(√πσt)
4.2. Branched filament
exp(− (δ−L cos(ω))2
erfc( L cos(ω)−δ
σ2
t
)
σt
)
.
Starting from Eq. 7 and averaging over graft point positions x0, we obtain
Pb(x2, x3) = (
(det(M))
1
2
π
) exp(ηiMijηj),
where
η = (cid:18)x2 − L1 cos(ω) − Lb cos(ω − γ)
x3 − (L1 + L3) cos(ω)
(cid:19)
and M is a 2 × 2-Matrix with components
M11 = −
lp
C
[K(L1 + L3)3 sin(ω)2 +
3
2
lp]
and
M12 = M21 =
3
2
[KL2
1(L1 +
L3) sin(ω)2 +
lp
C
3
KL1Lb(L1 + 2L3) sin(ω) sin(ω − γ)]
2
lp +
3
2
M22 = −lp
C
3KL2
b(3L1 + Lb) sin(ω − γ)2 +
[KL2
1Lb sin(ω) sin(ω − γ) + KL3
3
2
lp +
1 sin(ω)2],
4
3
C = K sin(ω)2[(L1 +
− 2(L2
+ L2
3)L2
1L2
L3)L3
3 sin(ω)2
1LbL3 sin(ω) sin(ω − γ)]
1 − 2L2
bK sin(ω)2 sin(ω − γ)2[L4
1LbL3 + 4L1L2
3(Lb + L3)]
L3)L2
3 sin(ω)2 − 2L1LbL3 sin(ω) sin(ω − γ)
1Lb + LbL3
3)
(L3
1 +
4
3
1
3
+ 6lp[(L1 +
+ 4L2
+ (L1 +
1
3
Lb)L2
b sin(ω − γ)2]
Mechanical properties of branched actin filaments
15
4.3. Parameter values
The independent parameters of the model are the F-actin persistence length lp, the
equilibrium distance of the filament tips from the graft plane δ0, the stiffness K of
the graft, and the branching angle γ which is fixed by the Arp2/3 complex at about
70 ◦ [38, 39, 40]. The branch point position is uniformly distributed and the tilt angle
ω according to the above distributions. The tilt angle and δ0 fix the mother filament
contour length as L = δ0/ cos(ω).
We explained in the Discussion, that a region juxtaposed to the leading edge
membrane with a width of 1.0-1.5 µm is similar to an experimental realization of the
network configuration in Fig. 1, since it is weakly cross-linked. Values for δ0 suggested by
these observations are in the range 1.0-1.5 µm. Branching will have an effect on network
properties in a configuration like shown in Fig. 1, if most filaments are branched, i.e., if
δ0 is larger than the average branch distance.
The average branch distance in steadily moving cells has been a matter of debate
in recent years. Svitkina et al. concluded 20-50 nm from early electron micrographs
of lamellipodia from Xenopus keratocytes and fibroblasts [15], and 50-200 nm from
another study in fibroblasts [41]. Later measurements substantially increased that
value. The average branch distance has been determined to be about 800 nm for
B16 melanoma cells and fish keratocytes in ref. [35]. Other studies provide number
densities of branch points per lamellipodium area for 3T3 fibroblasts. Calculating the
average branch distance from that branch point density implies assumptions on the F-
actin concentration. The number of filaments per micrometer lamellipodium width in
a distance of 0.1-1 µm from the leading edge is 170-190 [29]. Taking into account that
filament orientation is approximately uniformly distributed between 0◦ and 60◦ [27],
this density means 1.16(170-190) µm filament contour length per µm2 lamellipodium
area. The factor 1.16 arises from averaging over all tilt angles of the filaments. Yang
and Svitkina measured 277 branch points/µm2 [40] in the same sample, i.e. an average
branch point distance in terms of contour length between 700 nm and 800 nm. Small
et al. measured less than 225 branch points/µm2 [30], i.e. an average distance of more
than 860 nm.
The in vivo persistence length is not known. We can only conclude a reasonable
range from in vitro measurements. Results from fluctuation analysis and measurements
of network elastic properties yield values of the in vitro persistence length of 15-
18 µm [37, 42, 9, 43]. The filaments in these experiments were stabilized with phaloidin
which most likely increases the value of lp. Filaments labeled with rhodamine but not
stabilized with phaloidin exhibited values of lp between 9 µm and 13.5 µm [43]. The in
vivo persistence length might be even shorter, since cofilin can substantially reduce it
even down to 2.2 µm [44, 45].
Mechanical properties of branched actin filaments
16
ACKNOWLEDGMENTS
AZ acknowledges financial support by the DFG through Grant No. SFB 937/A1. MF
acknowledges financial support by the DFG through Grant No. FA350/11-1.
References
[1] D. Bray. Cell Movements. Garland, New York, 2nd edition, 2001.
[2] Hidekiand Yamaguchi, Jeffrey Wyckoff, and John Condeelis. Cell migration in tumors. Curr Opin
Cell Biol, 17:559 -- 564, Oct 2005.
[3] John Condeelis and Jeffrey W. Pollard. Macrophages: Obligate partners for tumor cell migration,
invasion, and metastasis. Cell, 124(2):263 -- 266, 2006.
[4] J. Victor Small, Theresia Stradal, Emmanuel Vignal, and Klemens Rottner. The lamellipodium:
where motility begins. Trends Cell Biol, 12(3):112 -- 120, 2002.
[5] T. M. Svitkina, A. B. Verkhovsky, K. M. McQuade, and G. G. Borisy. Analysis of the Actin-
Myosin II System in Fish Epidermal Keratocytes: Mechanism of Cell Body Translocation. J
Cell Biol, 139:397 -- 415, 1997.
[6] TD Pollard. The cytoskeleton, cellular motility and the reductionist agenda. Nature, 422:741 -- 745,
2003.
[7] Jingyuan Xu, Denis Wirtz, and Thomas D. Pollard. Dynamic cross-linking by α-actinin determines
the mechanical properties of actin filament networks. J Biol Chem, 273(16):9570 -- 9576, 1998.
[8] Alex Mogilner and George Oster. Force Generation by Actin Polymerization II: The Elastic
Ratchet and Tethered Filaments. Biophys J, 84(3):1591 -- 1605, 2003.
[9] M. L. Gardel, J. H. Shin, F. C. MacKintosh, L. Mahadevan, P. Matsudaira, and D. A. Weitz.
Elastic behavior of cross-linked and bundled actin networks. Science, 304(5675):1301 -- 1305,
2004.
[10] Sophie Bohnet, Revathi Ananthakrishnan, Alex Mogilner, Jean-Jacques Meister, and Alexander B.
Verkhovsky. Weak force stalls protrusion at the leading edge of the lamellipodium. Biophys J,
90(5):1810 -- 1820, 2006.
[11] Marcus Prass, Ken Jacobson, Alex Mogilner, and Manfred Radmacher. Direct measurement of
the lamellipodial protrusive force in a migrating cell. J Cell Biol, 174(6):767 -- 772, 2006.
[12] Fabian Heinemann, Holger Doschke, and Manfred Radmacher. Keratocyte lamellipodial
protrusion is characterized by a concave force-velocity relation. Biophys J, 100(6):1420 -- 1427,
2011.
[13] Juliane Zimmermann, Claudia Brunner, Mihaela Enculescu, Michael Goegler, Allen Ehrlicher,
Josef Kas, and Martin Falcke. Actin filament elasticity and retrograde flow shape the force-
velocity relation of motile cells. Biophys J, 102(2):287 -- 295, 2012.
[14] Fumihiko Nakamura, Eric Osborn, Paul A. Janmey, and Thomas P. Stossel. Comparison of
filamin a-induced cross-linking and arp2/3 complex-mediated branching on the mechanics of
actin filaments. J Biol Chem, 277(11):9148 -- 9154, 2002.
[15] Tatyana M. Svitkina and Gary G. Borisy. Arp2/3 complex and actin depolymerizing factor/cofilin
in dendritic organization and treadmilling of actin filament array in lamellipodia. J Cell Biol,
145(5):1009 -- 1026, 1999.
[16] A.B. Verkhovsky, T.M. Svitkina, and G.G. Borisy. Self-polarization and directional motility of
cytoplasm. Curr Biol, 9(1):11 -- 20, 1999.
[17] Pascal Vallotton, Gaudenz Danuser, Sophie Bohnet, Jean-Jacques Meister, and Alexander B.
Verkhovsky. Tracking Retrograde Flow in Keratocytes: News from the Front. Mol Biol Cell,
16(3):1223 -- 1231, 2005.
[18] Kinneret Keren, Zachary Pincus, Greg M. Allen, Erin L. Barnhart, Gerard Marriott, Alex
Mechanical properties of branched actin filaments
17
Mogilner, and Julie A. Theriot. Mechanism of shape determination in motile cells. Nature,
453(7194):475 -- 480, 2008.
[19] Juliane Zimmermann, Mihaela Enculescu, and Martin Falcke. Leading edge - gel coupling in
lamellipodium motion. Phys Rev E, 82(5):051925, Nov 2010.
[20] Boris Rubinstein, Maxime F. Fournier, Ken Jacobson, Alexander B. Verkhovsky, and Alex
Mogilner. Actin-myosin viscoelastic flow in the keratocyte lamellipod. Biophys J, 97(7):1853 --
1863, 2009.
[21] Mihaela Enculescu, Mohsen Sabouri-Ghomi, Gaudenz Danuser, and Martin Falcke. Modeling of
protrusion phenotypes driven by the actin-membrane interaction. Biophys J, 98(8):1571 -- 1581,
2010.
[22] Erin L. Barnhart, Kun-Chun Lee, Kinneret Keren, Alex Mogilner, and Julie A. Theriot. An
adhesion-dependent switch between mechanisms that determine motile cell shape. PLoS Biol,
9(5):e1001059, 05 2011.
[23] Danying Shao, Herbert Levine, and Wouter-Jan Rappel. Coupling actin flow, adhesion, and
morphology in a computational cell motility model. Proc Nat Acad Sci USA, 109(18):6851 --
6856, 2012.
[24] M. Falcke and J. Zimmermann. Polymerization, bending, tension: What happens at the leading
edge of motile cells? EPJ ST, 223(7):1353 -- 1372, 2014.
[25] P. Benetatos and E. Frey. Depinning of semiflexible polymers. Phys. Rev. E, 67:051108, May
2003.
[26] Azam Gholami, Jan Wilhelm, and Erwin Frey. Entropic forces generated by grafted semiflexible
polymers. Phys Rev E, 74(4):041803, Oct 2006.
[27] Stefan A. Koestler, Sonja Auinger, Marlene Vinzenz, Klemens Rottner, and J. Victor Small.
Differentially oriented populations of actin filaments generated in lamellipodia collaborate in
pushing and pausing at the cell front. Nat Cell Biol, 10(3):306 -- 313, 2008.
[28] Ivan V. Maly and Gary G. Borisy.
Self-organization of a propulsive actin network as an
evolutionary process. Proceedings of the National Academy of Sciences, 98(20):11324 -- 11329,
2001.
[29] W Urban, S Jacob, M Nemethova, GP Resch, and JV Small. Electron tomography reveals
unbranched networks of actin filaments in lamellipodia. Nat Cell Biol, 12:429 -- 435, 2010.
[30] J.V. Small, C. Winkler, M. Vinzenz, and C. Schmeiser. Reply: Visualizing branched actin filaments
in lamellipodia by electron tomography. Nat Cell Biol, 13(9):1013 -- 1014, 2011.
[31] Alexander B. Verkhovsky, Oleg Y. Chaga, Sebastien Schaub, Tatyana M. Svitkina, Jean-Jacques
Meister, and Gary G. Borisy. Orientational Order of the Lamellipodial Actin Network as
Demonstrated in Living Motile Cells. Mol Biol Cell, 14(11):4667 -- 4675, 2003.
[32] Julian Weichsel, Edit Urban, J. Victor Small, and Ulrich S. Schwarz. Reconstructing the
orientation distribution of actin filaments in the lamellipodium of migrating keratocytes from
electron microscopy tomography data. Cytometry Part A, 81A(6):496 -- 507, 2012.
[33] Christian H. Schreiber, Murray Stewart, and Thomas Duke. Simulation of cell motility that
reproduces the force-velocity relationship. Proc Nat Acad Sci USA, 107(20):9141 -- 9146, 2010.
[34] Dylan T. Burnette, Suliana Manley, Prabuddha Sengupta, Rachid Sougrat, Michael W. Davidson,
Bechara Kachar, and Jennifer Lippincott-Schwartz. A role for actin arcs in the leading-edge
advance of migrating cells. Nat Cell Biol, 13(4):371 -- 382, 2011.
[35] Marlene Vinzenz, Maria Nemethova, Florian Schur, Jan Mueller, Akihiro Narita, Edit Urban,
Christoph Winkler, Christian Schmeiser, Stefan A. Koestler, Klemens Rottner, Guenter P.
Resch, Yuichiro Maeda, and J. Victor Small. Actin branching in the initiation and maintenance
of lamellipodia. J Cell Sci, 125(11):2775 -- 2785, 2012.
[36] Sabine Petry, AaronC. Groen, Keisuke Ishihara, TimothyJ. Mitchison, and RonaldD. Vale.
Branching microtubule nucleation in xenopus egg extracts mediated by augmin and {TPX2}.
Cell, 152(4):768 -- 777, 2013.
[37] F Gittes, B Mickey, J Nettleton, and J Howard. Flexural rigidity of microtubules and actin
Mechanical properties of branched actin filaments
18
filaments measured from thermal fluctuations in shape. The Journal of Cell Biology, 120(4):923 --
934, 1993.
[38] R. Dyche Mullins, John A. Heuser, and Thomas D. Pollard. The interaction of Arp2/3 complex
with actin: Nucleation, high affinity pointed end capping, and formation of branching networks
of filaments. Proc Nat Acad Sci USA, 95(11):6181 -- 6186, 1998.
[39] Niels Volkmann, Kurt J. Amann, Svetla Stoilova-McPhie, Coumaran Egile, Dirk C. Winter,
Larnele Hazelwood, John E. Heuser, Rong Li, Thomas D. Pollard, and Dorit Hanein. Structure
of arp2/3 complex in its activated state and in actin filament branch junctions. Science,
293(5539):2456 -- 2459, 2001.
[40] Yang C. and Svitkina T. Visualizing branched actin filaments in lamellipodia by electron
tomography. Nat Cell Biol, 13(9):1012 -- 1013, 2011.
[41] James E. Bear, Tatyana M. Svitkina, Matthias Krause, Dorothy A. Schafer, Joseph J. Loureiro,
Geraldine A. Strasser, Ivan V. Maly, Oleg Y. Chaga, John A. Cooper, Gary G. Borisy, and
Frank B. Gertler. Antagonism between ena/vasp proteins and actin filament capping regulates
fibroblast motility. Cell, 109(4):509 -- 521, 2002.
[42] Loıc Le Goff, Oskar Hallatschek, Erwin Frey, and Fran¸cois Amblard. Tracer studies on f-actin
fluctuations. Phys Rev Lett, 89(25):258101, Dec 2002.
[43] H Isambert, P Venier, AC Maggs, A Fattoum, R Kassab, D Pantaloni, and MF Carlier. Flexibility
of actin filaments derived from thermal fluctuations. effect of bound nucleotide, phalloidin, and
muscle regulatory proteins. J Biol Chem, 270(19):11437 -- 11444, 1995.
[44] Brannon R. McCullough, Laurent Blanchoin, Jean-Louis Martiel, and Enrique M. De La Cruz.
Cofilin increases the bending flexibility of actin filaments: Implications for severing and cell
mechanics. J Mol Biol, 381(3):550 -- 558, 2008.
[45] Jim Pfaendtner, Enrique M. De La Cruz, and Gregory A. Voth. Actin filament remodeling by
actin depolymerization factor/cofilin. Proc Nat Acad Sci USA, 107(16):7299 -- 7304, 2010.
|
1802.05769 | 3 | 1802 | 2019-01-21T20:23:23 | The polar clasps of a bank vole PrP(168--176) prion protofibril revisiting | [
"physics.bio-ph",
"q-bio.BM"
] | On 2018-01-17 two electron crystallography structures (with PDB entries 6AXZ, 6BTK) on a prion protofibril of bank vole PrP(168-176) (a segment in the PrP $\beta$2-$\alpha$2 loop) were released into the PDB Bank. The paper published by [Nat Struct Mol Biol 25(2):131-134 (2018)] reports some polar clasps for these two crystal structures, and "an intersheet hydrogen bond between Tyr169 and the backbone carbonyl of Asn171 on an opposing strand." - this hydrogen bond is not between the neighbouring Chain B and Chain A directly. In addition, by revisiting the polar clasps, we found another two hydrogen bonds ([email protected]@OE1, [email protected]@N) between the strand A of one sheet and the opposing strand B of the mating sheet. For the neighbouring two single $\beta$-sheets AB, the two new hydrogen bonds are completely different from the experimental one (an intersheet hydrogen bond between Tyr169 and the backbone carbonyl of Asn171 on an opposing strand) in [Nat Struct Mol Biol 25(2):131-134 (2018)]. | physics.bio-ph | physics |
The polar clasps of a bank vole PrP(168 -- 176) prion protofib-
ril revisiting
Jiapu Zhang
Centre of Informatics and Applied Optimisation, The Federation University Australia, Mount
Helen Campus, Mount Helen, Ballarat, Victoria 3353, Australia;
*Tel: +61-3-5327 6335; Email: [email protected]
Abstract: On 2018-01-17 two electron crystallography structures (with PDB entries
6AXZ, 6BTK) on a prion protofibril of bank vole PrP(168-176) (a segment in the PrP
β2-α2 loop) were released into the PDB Bank. The paper published by [Nat Struct Mol
Biol 25(2):131-134 (2018)] reports some polar clasps for these two crystal structures, and
"an intersheet hydrogen bond between Tyr169 and the backbone carbonyl of Asn171
on an opposing strand." - this hydrogen bond is not between the neighbouring Chain B
and Chain A directly. In addition, by revisiting the polar clasps, we found another two
hydrogen bonds ([email protected]@OE1, [email protected]@N) between
the strand A of one sheet and the opposing strand B of the mating sheet. For the
neighbouring two single β-sheets AB, the two new hydrogen bonds are completely
different from the experimental one (an intersheet hydrogen bond between Tyr169 and
the backbone carbonyl of Asn171 on an opposing strand) in [Nat Struct Mol Biol
25(2):131-134 (2018)].
Key words: PrP structured region · Fibril formation peptides · The 'polar clasps' ·
Mathematical formulas and optimizations · The hydrogen bonds
1
Introduction
Prion diseases were caused by the conformational change from normal cellular prion
protein PrPC into diseased infectious prion PrPSc. Numerous structures of PrPC are
known and have been deposited into the Protein Data Bank (rcsb.org) (PDB). How-
ever, the atomic structure of the infectious, protease-resistant, β-sheet-rich and fibrillar
mammalian PrPSc remains unknown, being due to the unstable, noncrystalline and in-
soluble nature of the fibrils. Only during recent a decade the lab of Eisenberg DS
at University of California Los Angeles and other labs are producing the atomic struc-
tures of some segments of PrPSc. On 2018-01-17 two electron crystallography structures
(with PDB entries 6AXZ, 6BTK) on a prion protofibril of bank vole PrP(168 -- 176) (a
segment in the PrP β2-α2 loop) were released into the PDB Bank. The structures
belong to the ordinary Class 2 steric zipper with β-strands parallel and sheets pair
front to back [1], but they have an outstanding hallmark: within a sheet there are very
rich hydrogen bonds (HBs) - even constructing several 'circles' by polar residues Gln,
Asn and Tyr of the two neighboring β-strands of a sheet; and between two neighbor-
ing β-sheets (if each is with one and only one β-strand) there are two additional HBs
(B.Asn171@H -- A.Gln172@OE1, B.Tyr169@OH -- A.Gln172@N) too - besides the one of
[2]. The authors named this character as 'polar clasps'. This brief paper is revisiting
the 'polar clasps' from the points of view in crystal mathematical formulas and math-
ematical optimization. Detailed HBs will be presented and illuminated for the 'polar
1
clasps'.
First we introduce the definition of 'polar clasps' in [2]. 'Polar clasp' is a motif whose
neighboring polar ladders in the PrPSc are linked by HBs within a strand. Stacks of
aromatic residues Phe175 and Tyr169 will shield these clasps at the core of PrPSc.
In [3], the author showed us in Year 2016 a table listing all the PrP segments with
cross-β structures and each model is with its mathematical formulas and a color photo.
Now we insert two elements ((21) -- (22)) of bank vole (Myodes glareolus) PrP(168 -- 176)
of Year 2018 into the table and produce a new table, Tab. 1. We organize this paper as
follows. Firstly, we present the approaches and tools that will be used in this paper and
our major discovery about (molecular modelling) optimization skills will be given in
the Methods Section. In the Results and Discussion Section, for models (21) -- (22), the
first Subsection will present mathematical formulas accompanied by graphs, then the
second Subsection will revisit HBs of the 'polar clasps'. Lastly, the Summary Section
gives four concluding remarks.
Table 1: The cross-β structures known in the PDB Bank of PrP segments:
Class of the cross-β
PDB ID
PrP segment
Species
(01) PrP(126 -- 131)
(02)
(03)
(04) PrP(126 -- 132)
(05)
(06)
(07) PrP(127 -- 132)
(08)
(09)
(10)
(11)
(12)
(13) PrP(127 -- 133)
(14)
(15)
(16) PrP(137 -- 142)
(17) PrP(137 -- 143)
(18) PrP(138 -- 143)
(19)
(20) PrP(142 -- 166)
(21) PrP(168 -- 176)
(22)
(23) PrP(170 -- 175)
(24)
(25) PrP(177 -- 182, 211 -- 216)
(26)
human
human-V129
human-L129
human-L129
human
human-V129
human
human-V129
human
human-V129
human
human-V129
human
human-V129
human-L129
mouse
mouse
Syrian hamster
human
sheep
bank vole
bank vole
human
elk
human
human
4TUT
4UBY
4UBZ
4W5L
4W5M
4W5P
4WBU
4WBV
3MD4
3MD5
3NHC
3NHD
4W5Y
4W67
4W71
3NVG
3NVH
3NVE
3NVF
1G04
6AXZ
6BTK
2OL9
3FVA
4E1I
4E1H
Class 7[1, 4]
Class 8[1, 4]
Class 8[1, 4]
Class 8[1, 4]
Class 8[1, 4]
Class 8[1, 4]
Class 8[1, 4]
Class 8[1, 4]
antiparallel (P 21 21 21)
parallel (P 1 21 1)
Class 8[1]
Class 8[1]
Class 6[1, 4]
Class 6[1, 4]
Class 6[1, 4]
Class 1[1]
Class 1[1]
Class 6[1]
Class 1[1]
β-hairpin[5]
Class 2[2, 1]
Class 2[2, 1]
Class 2[1]
Class 1[1]
β-prism I fold[6] (P 21 21 21)
β-prism I fold[6] (P 21 21 21)
2
2 Methods - The Molecular Modelling Approaches and
Tools
The molecular modelling approaches used in the paper are the deep analyses for the
files from the PDB library, optimization theory and the hybrid of optimization meth-
ods. The tools used are JSmol (http://www.jmol.org) - an open-source JavaScript
computer software for molecular modelling chemical structures in 3-dimensions, VMD
[7] (where the VMD is the most recent version VMD1.9.3Win32cuda downloaded from
http://www.ks.uiuc.edu/Research/vmd/) - a molecular visualization program for dis-
playing, animating, and analysing large biomolecular systems using 3-D graphics and
built-in scripting, and the Swiss-PdbViewer 4.1.0 (https://spdbv.vital-it.ch/) [8, 9] - a
free package for comparative protein modelling.
Optimization plays important roles in molecular modelling, and molecular simula-
tions of molecular dynamics and/or quantum mechanics/molecular mechanics. Large-
scale optimization computations have drawn considerable attention. The existing lo-
cal/global optimization techniques effectively solve many problems when the number of
variables is not very large and, as a rule, fail to solve many large-scale problems. The
study of new algorithms which allow one to solve large-scale optimization problem is
very important. One technique is to use hybrid of local/global and local/global search
algorithms. This paper advocates the use of hybrid algorithmic approaches [10]. It is
clearly possible that different optimization strategies would have made different results.
In the second subsection of the Results and Discussion section we will use the hybrid
of local search optimization methods Steepest Descent (SD) method + Conjugate Gra-
dient (CG) method + Steepest Descent (SD) method when using Swiss-PdbViewer
4.1.0.
3 Results and discussion
3.1 Mathematical formulas and graphs
In this subsection, we will give the mathematical formulas for the fibrils of bank vole
PrP(168 -- 176) with PDB entries 6AXZ and 6BTK respectively, then we will illuminate
their respective photos. The mathematical formulas can be gotten after reading the
PDB files at websites:
https://files.rcsb.org/view/6AXZ.pdb, https://files.rcsb.org/view/6BTK.pdb,
where the coordinates of the A Chain were given and we were told clearly "APPLY
THE FOLLOWING TO CHAINS: A" to get other Chains. Their photos can be gotten
from the websites https://www.rcsb.org/3d-view/6AXZ/1, https://www.rcsb.org/3d-
view/6BTK/1 respectively by selecting the viewer JSmol (JavaScript).
3
3.1.1 PDB entry 6AXZ -- bank vole PrP(168 -- 176)
Model (21): In Fig. 1 we see that B chain (i.e. β-sheet 2) of 6AXZ.pdb can be obtained
from A chain (i.e. β-sheet 1) by
B = A +
−2.18510
10.10648
0
,
and other chains can be got by
,
0
0
4.94
C(D) = A(B) +
E(F ) = A(B) +
2 ∗ 4.94
G(H) = A(B) +
I(J) = A(B) +
3 ∗ 4.94
4 ∗ 4.94
0
0
0
0
0
0
.
,
,
(1)
(2)
(3)
(4)
(5)
3.1.2 PDB entry 6BTK -- bank vole PrP(168 -- 176)
Model (22): In Fig. 2 we see that B chain (i.e. β-sheet 2) of 6BTK.pdb can be obtained
from A chain (i.e. β-sheet 1) by
B = A +
−2.04266
9.89843
0
,
and other chains can be got by
,
0
0
4.874
C(D) = A(B) +
E(F ) = A(B) +
2 ∗ 4.874
G(H) = A(B) +
I(J) = A(B) +
4 ∗ 4.874
3 ∗ 4.874
0
0
0
0
0
0
.
,
,
4
(6)
(7)
(8)
(9)
(10)
Figure 1: Protein fibril structure of QYNNQNNFV segment 168 -- 176 from bank vole prion
(PDB entry 6AXZ). The dashed lines denote the HBs. A, B, ..., I, J denote the 10 chains
of the fibril.
3.2 Revisiting the HBs
We used the above mathematical formulas and the Figs. 1-2 to get (I) AC-Chains and
(II) AB-Chains for 6AXZ.pdb and 6BTK.pdb respectively, and (III) ABCD-Chains
for 6AXZ.pdb. We used VMD to visualize the HBs of (I) -- (III) for each pdb but
we cannot find HBs yet (where the HB parameters are the default ones of VMD:
Donor-Acceptor Distance Cutoff 3.0 A, Angle cutoff 20 degrees). Thus we used Swiss-
PdbViewer 4.1.0 (the file SPDBV 4.10 PC.zip downloaded from https://spdbv.vital-
it.ch/) to optimize (I) -- (III) (Tab. 2), where we found the hydrogen atoms were added
after the optimization by the Swiss-PdbViewer.
The Swiss-PdbViewer is the most recent version Swiss-PdbViewer 4.1.0 and we set
the Energy Minimization Preferences as follows: 100 Steps of Steepest Descent (SD)
method, then 100 Steps of Conjugate Gradient (CG) method, and at last 50 Steps of
5
Figure 2: Protein fibril structure of QYNNQNNFV segment 168 -- 176 from bank vole prion
(PDB entry 6BTK). The dashed lines denote the HBs. A, B, ..., I, J denote the 10 chains
of the fibril.
SD method to slightly optimize the model - i.e. here we used the hybrids of SD-CG-
SD optimization methods. The optimization considered the Bonds, Angles, Torsions,
Improper, Non-bonded, Electrostatic, and we selected the Cutoff as 12.000 A. The
stopping criteria when delta E between two steps is below 0.050 KJ/mol and when
Force acting on any atom is below 10.000 are selected. The computations were done in
vacuo with the GROMOS96 43B parameters set, without reaction field, and the energy
computations were done with the GROMOS96 implementation of Swiss-PdbViewer
4.1.0.
After the optimization, we got the HBs network for (I) -- (III) respectively (Tabs.
3-5), where the HB detection threshold parameters are the default ones of Swiss-
PdbViewer 4.1.0: when Hydrogens are present, Min cutoff distance 1.200 A, Max cutoff
distance 2.760 + 0.050 A, Min Cutoff Angle 120.000 degrees; and when Hydrogens are
not present, Min cutoff distance 2.195 A, Max cutoff distance 3.300 + 0.050 A, Min
6
Table 2: The energy decreases during the optimization:
PrP segment
(21) PrP(168 -- 176)
Species
bank vole
PDB ID Chains Energy (KJ/mol) decreases
6AXZ
(22)
bank vole
6BTK
AC
-2250.996 → -2264.804 → -2268.384
ACE
-3617.681 → -3635.713 → -3646.515
ABCD -4706.212 → -4715.012 → -4726.107
-1885.567 → -1901.603 → -1904.400
-2268.958 → -2269.830 → -2274.719
-3664.371 → -3674.300 → -3678.038
-1916.021 → -1919.213 → -1922.448
AB
AC
ACE
AB
Cutoff Angle 90.000 degrees.
(I) AC-Chains. There are 20 HBs for the optimized AC Chains of 6AXZ and 6BTK
respectively (Tab. 3 and Fig. 3). Here we found 14 interchain HBs, including the 6
interchain HBs in Fig. 2 of [2] labeled by blue arrows.
Table 3: The 20 HBs between the optimized AC Chains or within its each Chain for each
pdb model:
in C Chain
in A Chain
Asn174@OD1 -- Gln172@H
Asn174@OD1 -- Gln172@H
ASN171@OD1 -- Asn173@H Asn171@OD1 -- Asn173@H
Asn170@H -- Gln168@OE1
Asn170@H -- Gln168@OE1
between C -- A Chains
Val176@H -- Phe175@O
Asn174@O -- Phe175@H
Asn174@H -- Asn173@O
Asn174@H -- Asn174@OD1
Asn173@OD1 -- Asn173@H
Gln172@H -- Asn171@O
Gln172@OE1 -- Gln172@H
Gln172@O -- Asn173@H
Asn171@OD1 -- Asn171@H
Asn170@O -- Asn171@H
Asn170@H -- Asn170@OD1
Asn170@H -- Tyr169@O
Gln168@O -- Tyr169@H
Gln168@OE1 -- Gln168@H
Figure 3: The optimized AC Chains in the fibril structures of QYNNQNNFV segment
168 -- 176 from bank vole prion (left: 6AXZ, right: 6BTK). The dashed lines denote the
HBs. The photos were produced by Swiss-PdbViewer 4.1.0. The green coloured A and C
denote the A Chain and C Chain respectively.
(II) AB-Chains. There are 7 HBs for the optimized AB Chains of 6AXZ and
7
6BTK respectively (Tab. 4 and Fig. 4). Here the single intersheet HB Asn171@H --
Gln172@OE1 is different from the one in Fig. 2 of [2] labeled by a red arrow (a note:
if we set the Max distance 3.4 A when Hydrogens are not present, we can find a HB
B.Tyr169@OH -- A.Gln172@N with distance 3.44 A (Fig. 5) where the coordinate of
B.Tyr169@OH is (-3.491, 1.826, 6.481), the coordinate of A.Gln172@N is (-1.581, -
0.424, 8.254), and B Chain is got from A Chain by mathematical formula (6). We
should notice here the AB Chains are not optimized yet). The intrachain three HBs
labeled by orange arrows in Fig. 2 of [2] are found here (see Tab. 4 and Fig. 4).
Table 4: The 7 HBs between the optimized AB Chains or within its each Chain for each
pdb model:
between B -- A Chains
Asn171@H -- Gln172@OE1
in A Chain
in B Chain
Asn174@OD1 -- Gln172@H Asn174@OD1 -- Gln172@H
Asn173@H -- Asn171@OD1 Asn173@H -- Asn171@OD1
Asn170@H -- Gln168@OE1 Asn170@H -- Gln168@OE1
Figure 4: The optimized AB Chains in the fibril structures of QYNNQNNFV segment
168 -- 176 from bank vole prion (left: 6AXZ, right: 6BTK). The dashed lines denote the
HBs. The photos were produced by Swiss-PdbViewer 4.1.0. The green coloured A and B
denote the A Chain and B Chain respectively.
The AB sheets stack in a parallel face-to-back configuration, the convex face of sheet
A nestles against the concave back of its neighboring sheet B, with a high degree of
surface complementarity (Fig. 4). Seeing Fig. 6, in AB Chains, the most colored in pink
are the polar residues Gln168 -- Asn174, another color of residues are the hydrophobic
residues Phe175 and Val176.
(III) ABCD-Chains. In ABCD Chains, 28.9% of the atoms are hydrogen. There
are a lot of HBs for the optimized ABCD Chains of 6AXZ model (see Fig. 7, and
the supplementary Tab. 2 of [2] where B is the C in here and C is the B here),
including the Asn171@H -- Gln172@OE1 in Tab. 4 linking B and A Chains, the HBs
Asn171@O -- Tyr169@H and Asn174@H -- Asn173@OD1 linking A and D Chains, and
Gln172@OE1 -- Asn171@H linking C and D Chains (Tab. 5 and Fig. 7).
By Figs. 7-8 and Tab. 5, we can confirm the "intersheet hydrogen bond between
Tyr169 (of D Chain) and the backbone carbonyl of Asn171 (of A Chain) on an opposing
8
Figure 5: The HB B.Tyr169@OH -- A.Gln172@N between AB Chains (not optimized yet,
directly got by mathematical formula (6)) in the fibril structures of QYNNQNNFV segment
168 -- 176 from bank vole prion (6BTK.pdb). The dashed lines denote the HBs. The photo
was produced by Swiss-PdbViewer 4.1.0.
Figure 6: The distributions of residues of AB Chains in the fibril structures of QYN-
NQNNFV segment 168 -- 176 from bank vole prion (PDB entry 6AXZ). Pink: polar residues
Gln168 -- Asn174, gray: hydrophobic residues Phe175 -- Val176.
Table 5: The 4 special HBs between the optimized ABCD Chains for the 6AXZ.pdb model:
The pair of Chains HBs
BA
AD
CD
Asn171@H -- Gln172@OE1
Asn171@O -- Tyr169@H, Asn174@H -- Asn173@OD1
Gln172@OE1 -- Asn171@H
strand" of [2], bearing in mind that
D = A +
−2.18510 + 4.94
10.10648
0
.
(11)
9
Figure 7: The optimized ABCD Chains in the fibril structures of QYNNQNNFV segment
168 -- 176 from bank vole prion (PDB entry 6AXZ). The dashed lines denote the HBs. The
photos were produced by Swiss-PdbViewer 4.1.0.
Figure 8: The HB A.Asn171@O -- D.Tyr169@H between the optimized AD Chains in the
fibril structures of QYNNQNNFV segment 168 -- 176 from bank vole prion (6AXZ.pdb).
The dashed lines denote the HB. The photo was offered by Rodriguez JA (University of
California Los Angeles).
10
At the end of this subsection, we revisit the HBs of the optimized ACE Chains.
We illuminate them in Fig. 9, where there are some 'polar clasps' (as shown in the
supplementary Fig. 7 of [2]).
Figure 9: The optimized ACE Chains in the fibril structures of QYNNQNNFV segment
168 -- 176 from bank vole prion (left: 6AXZ, right: 6BTK). The dashed lines denote the
HBs. The photos were produced by Swiss-PdbViewer 4.1.0. The green coloured A, C, E
denote the A Chain, C Chain and E Chain respectively.
4 Summary
This short paper revisited the 'polar clasps' and presented detailed HBs of the 'polar
clasps'. We found out more HBs for the neighboring two β-strands, for the block
of two β-sheets - each sheet being with two β-strands, and the most important, for
the neighboring two single β-sheets AB, two new HBs B.Asn171@H -- A.Gln172@OE1,
B.Tyr169@OH -- A.Gln172@N were found out, which is completely different from the
experimental one (an intersheet HB between Tyr169 and the backbone carbonyl of
Asn171 on an opposing strand) in [2].
The structures of some segments of PrPSc have been determined, including the most
recent bank vole PrPSc(168 -- 176) one of [2]. In Fig. 16.3 of [11], we have predicted that
some segments PrP(126 -- 133), PrP(137 -- 143), PrP(168 -- 176), PrP(170 -- 175), PrP(177 --
182), PrP(211 -- 216) should have the amyloid fibril forming property.
This brief paper also showed us the importance of optimization. Before optimiza-
tion, we could not find out any HBs (where the distance and angle cutoffs are the
default values of VMD) from the structures calculated by mathematical formulas given
according to crystal informatics in the PDB files. However, after optimization, all the
HBs were revealed. The correct hybrid skill of optimization algorithms is also worth
having a mention here. In the short comments, the author revisited two recently re-
vealed Crystal structures of PrP segments with "polar collapse" and found extra HBs in
the network after optimization. The manuscript is clear written. Here we discuss more
on comparing the structures in the PDB library and those after optimizations. How far
away the current structures deposited in the PDB library from the refined structures
are listed as follows. For 6AXZ.pdb, its RMSD [12] values are 0.077589, 0.058435,
0.075521, and 0.098501 angstroms for ABCD Chains, AB Chains, AC Chains, and
ACE Chains respectively. For 6BTK.pdb, its RMSD values are 0.056441, 0.045499,
and 0.075953 angstroms for AB Chains, AC Chains, and ACE Chains respectively.
These RMSD values could further highlight the importance/necessity of refinement in
the protein structure study.
11
In this paper, we revisited the 'polar clasps' in the structures of PrP and presented
detailed HBs of the "polar clasps" after optimization.
It is possible that different
optimization strategies would have made different results. However, if the local searches
in Swiss-PdbViewer 4.1.0 were carried out thoroughly, the optimization results should
be the same ones.
Acknowledgments
This research was supported by a Melbourne Bioinformatics grant numbered FED0001
on its Peak Computing Facility at the University of Melbourne, an initiative of the
Victorian Government (Australia). The author thanks Rodriguez JA (University of
California Los Angeles) for his helps to understand their paper [2] deeply. The author
is also grateful to some comments from reviewers, which have improved this paper
greatly.
References
[1] Sawaya MR, Sambashivan S, Nelson R, Ivanova MI, Sievers SA, Apostol MI,
Thompson MJ, Balbirnie M, Wiltzius JJ, McFarlane HT, Madsen A, Riekel C,
Eisenberg D (2007) Atomic structures of amyloid cross-β spines reveal varied steric
zippers. Nature 447(7143):453 -- 457
[2] Gallagher-Jones M, Glynn C, Boyer DR, Martynowycz MW, Hernandez E, Miao
J, Zee CT, Novikova IV, Goldschmidt L, McFarlane HT, Helguera GF, Evans JE,
Sawaya MR, Cascio D, Eisenberg DS, Gonen T, Rodriguez JA (2018) Sub-ngstrm
cryo-EM structure of a prion protofibril reveals a polar clasp. Nat Struct Mol Biol
25(2):131 -- 134
[3] Zhang JP (2016) Mathematical formulas for prion all cross-β structures listed in
the Protein Data Bank. Med Chem 6(3):179 -- 188
[4] Yu L, Lee SJ, Yee VC (2015) Crystal structures of polymorphic prion protein β1
peptides reveal variable steric zipper conformations. Biochemistry 54(23):3640 --
3648
[5] Kozin SA, Bertho G, Mazur AK, Rabesona H, Girault JP, Haertl T, Takahashi
M, Debey P, Hoa GH (2001) Sheep prion protein synthetic peptide spanning helix
1 and beta-strand 2 (residues 142-166) shows beta-hairpin structure in solution. J
Biol Chem 276(49):46364 -- 46370
[6] Apostol MI, Perry K, Surewicz WK (2013) Crystal structure of a human prion
for oligomer formation. J Am Chem Soc
protein fragment reveals a motif
135(28):10202 -- 10205
[7] Humphrey W, Dalke A, Schulten K (1996) VMD: visual molecular dynamics. J
Mol Graph 14(1):33 -- 38
12
[8] Guex N, Peitsch MC, Schwede T (2009) Automated comparative protein structure
modeling with SWISS-MODEL and Swiss-PdbViewer: a historical perspective.
Electrophoresis 30 Suppl 1:S162 -- 173
[9] Guex N, Peitsch MC (1997) SWISS-MODEL and the Swiss-PdbViewer: An envi-
ronment for comparative protein modeling. Electrophoresis 18: 2714 -- 2723
[10] Zhang
JP (2011) Derivative-free
hybrid methods
2010.
global
opti-
in
LAMBERT Aca-
(Available
online
and
mization
applications
demic Publishing, Gemany.
http://researchonline.federation.edu.au/vital/access/HandleResolver/1959.17/137122)
ISBN 978 -- 3 -- 8454 -- 3580 -- 0
in December
[11] Zhang JP (2015) Molecular Structures and Structural Dynamics of Prion Proteins
and Prions: Mechanism Underlying the Resistance to Prion Diseases. Springer,
Dordrecht. ISBN:978 -- 94 -- 017 -- 7317 -- 1
[12] Zhang JP (2018) Molecular Dynamics Analyses of Prion Protein Structures (1st ed.
2018 Edition): The Resistance to Prion Diseases Down Under, Springer, Singapore.
ISBN 978 -- 981 -- 10 -- 8814 -- 8
13
|
1804.09711 | 2 | 1804 | 2018-06-12T16:01:53 | Delocalized excitons in natural light harvesting complexes | [
"physics.bio-ph",
"cond-mat.soft",
"quant-ph"
] | Natural organisms such as photosynthetic bacteria, algae, and plants employ complex molecular machinery to convert solar energy into biochemical fuel. An important common feature shared by most of these photosynthetic organisms is that they capture photons in the form of excitons typically delocalized over a few to tens of pigment molecules embedded in protein environments of light harvesting complexes (LHCs). Delocalized excitons created in such LHCs remain well protected despite being swayed by environmental fluctuations, and are delivered successfully to their destinations over hundred nanometer length scale distances in about hundred picosecond time scales. Decades of experimental and theoretical investigation have produced a large body of information offering insights into major structural, energetic, and dynamical features contributing to LHCs' extraordinary capability to harness photons using delocalized excitons. The objective of this review is (i) to provide a comprehensive account of major theoretical, computational, and spectroscopic advances that have contributed to this body of knowledge, and (ii) to clarify the issues concerning the role of delocalized excitons in achieving efficient energy transport mechanisms. The focus of this review is on three representative systems, Fenna-Matthews-Olson complex of green sulfur bacteria, light harvesting 2 complex of purple bacteria, and phycobiliproteins of cryptophyte algae. Although we offer more in-depth and detailed description of theoretical and computational aspects, major experimental results and their implications are also assessed in the context of achieving excellent light harvesting functionality. Future theoretical and experimental challenges to be addressed in gaining better understanding and utilization of delocalized excitons are also discussed. | physics.bio-ph | physics | Delocalized excitons in natural light harvesting complexes
Seogjoo J. Jang
Department of Chemistry and Biochemistry,
Queens College, City University of New York,
65-30 Kissena Boulevard, Queens,
NY 11367 & PhD programs in Chemistry and Physics,
and Initiative for Theoretical Sciences,
Graduate Center,
City University of New York,
365 Fifth Avenue, New York,
NY 10016
Benedetta Mennucci
Department of Chemistry,
University of Pisa,
via G. Moruzzi 13, 56124 Pisa,
Italy
(Dated: Accepted for publication in the Reviews of Modern Physics on March 26, 2018; Revised on June 8, 2018)
Natural organisms such as photosynthetic bacteria, algae, and plants employ complex
molecular machinery to convert solar energy into biochemical fuel. An important com-
mon feature shared by most of these photosynthetic organisms is that they capture pho-
tons in the form of excitons typically delocalized over a few to tens of pigment molecules
embedded in protein environments of light harvesting complexes (LHCs). Delocalized
excitons created in such LHCs remain well protected despite being swayed by environ-
mental fluctuations, and are delivered successfully to their destinations over hundred
nanometer length scale distances in about hundred picosecond time scales. Decades of
experimental and theoretical investigation have produced a large body of information
offering insights into major structural, energetic, and dynamical features contributing
to LHCs' extraordinary capability to harness photons using delocalized excitons. The
objective of this review is (i) to provide a comprehensive account of major theoretical,
computational, and spectroscopic advances that have contributed to this body of knowl-
edge, and (ii) to clarify the issues concerning the role of delocalized excitons in achieving
efficient energy transport mechanisms. The focus of this review is on three representa-
tive systems, Fenna-Matthews-Olson complex of green sulfur bacteria, light harvesting
2 complex of purple bacteria, and phycobiliproteins of cryptophyte algae. Although we
offer more in-depth and detailed description of theoretical and computational aspects,
major experimental results and their implications are also assessed in the context of
achieving excellent light harvesting functionality. Future theoretical and experimental
challenges to be addressed in gaining better understanding and utilization of delocalized
excitons are also discussed.
I. INTRODUCTION
The energy for life on earth begins with harvesting
of photons from sunlight. To perform this quantum
mechanical process, photosynthetic organisms have de-
veloped a wide range of highly tuned forms of light
harvesting complexes (LHCs) (van Amerongen et al.,
2000; Blankenship, 2014; Cogdell et al., 2006; Hu et al.,
2002), while utilizing surprisingly few kinds of pigment
molecules (Blankenship, 2014; Croce and van Ameron-
gen, 2014) as primary units that absorb photons and
create electronic excitations (excitons) (Agranovich and
Galanin, 1982; Davydov, 1971). Most LHCs consist of
about 10 − 100 pigment molecules held by protein scaf-
folds, with typical inter-pigment distances (cid:38) 1nm, and
harness excitons that are typically delocalized over a few
to tens of such pigment molecules. In general, the light
harvesting unit of a specific photosynthetic organism is
formed by aggregates or super-complexes of those LHCs.
Each LHC exhibits unique structural and energetic fea-
tures realized through specific arrangement of pigment
molecules with finely tuned excitation energies (Blanken-
ship, 2014), which in turn determine the energetics and
the dynamics of delocalized excitons. These are believed
to embody certain design principles that have been de-
veloped through long evolutionary adaptation processes,
and ensure optimal light harvesting capability for given
living environments and specific light conditions. Figure
1 provides a simple overview of the evolutionary tree of
photosynthetic organisms and major LHCs.
To what extent do such design principles owe their suc-
cess to quantum effects associated with the delocalized
nature of excitons? Answering this question might have
been the underlying motivation for decades of research on
8
1
0
2
n
u
J
2
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
1
1
7
9
0
.
4
0
8
1
:
v
i
X
r
a
LHCs. However, only during the past decade, this ques-
tion has become much more explicit, attracting numer-
ous experimental and theoretical studies. While there
have been excellent review articles on various aspects of
LHCs (Cheng and Fleming, 2009; Chenu and Scholes,
2015; Cogdell et al., 2006; Curutchet and Mennucci, 2017;
Hu et al., 2002; Jang and Cheng, 2013; Jankowiak et al.,
2011; Kondo et al., 2017; Konig and Neugebauer, 2012;
Lambert et al., 2013; Lee et al., 2016; Levi et al., 2015;
Mirkovic et al., 2017; Olaya-Castro et al., 2012; Olaya-
Castro and Scholes, 2011; Pach´on and Brumer, 2012;
Renger, 2009; Renger et al., 2001; Renger and Muh, 2013;
Schroter et al., 2015), a comprehensive review for the
general physics community, which encompasses all of the
energetic, dynamic, and spectroscopic information, seems
currently lacking to the best of our knowledge. Our ob-
jective is to offer such a review with particular focus on
quantum mechanical aspects of delocalized excitons in
LHCs.
Excitons created in most LHCs are generally viewed as
Frenkel type (Davydov, 1971; Frenkel, 1931), for which
the Hamiltonian representing a single exciton can be ex-
pressed as
He =
Ejsj(cid:105)(cid:104)sj +
Jjksj(cid:105)(cid:104)sk ,
(1)
(cid:88)
(cid:88)
k(cid:54)=j
j
(cid:88)
j
where sj(cid:105) represents an exciton localized at site j,
namely, an individual pigment or chromophore in LHCs,
with Ej as the corresponding energy. The term Jjk rep-
resents the electronic coupling between site local exciton
states sj(cid:105) and sk(cid:105). In general, this is a sum of Coulomb
and exchange interactions between the two site excita-
tion states. Another commonly accepted assumption of
Frenkel-type exciton, although not essential, is that sj(cid:105)'s
form an orthogonal basis. This assumption can be jus-
tified for most LHCs, where different pigment molecules
are separated well enough to have negligible overlap inte-
grals between respective electronic excited states. Under
these assumptions, the exciton Hamiltonian can be diag-
onalized easily as
He =
Ejϕj(cid:105)(cid:104)ϕj ,
(2)
j
where Ej is the eigenvalue and ϕj(cid:105) is the corresponding
eigenstate. This is related to the site excitation state
through a unitary transformation with matrix element,
Ujk = (cid:104)skϕj(cid:105), as follows:
ϕj(cid:105) =
Ujksk(cid:105) .
(cid:88)
(3)
k
(cid:88)
2
(a)
(b)
FIG. 1 (a) A simple evolutionary tree of photosynthetic or-
ganisms (provided by Gregory D. Scholes) (b) Major light
harvesting antennas and their spectral regions adapted with
permission from (Scholes, 2010). Copyright 2010 American
Chemical Society.
the magnitudes of Jjk's relative to those of δEjk's, where
δEjk = Ej − Ek. However, as in most condensed phase
molecular systems, Eq. (1) or (2) alone is severely lim-
ited in representing the full quantum nature of excitons
in LHCs. Other factors such as molecular vibrations and
polarization response of protein environments have to be
taken into consideration as well. This requires consider-
ation of additional Hamiltonian terms.
For Frenkel excitons, delocalization occurs only
through a quantum mechanical superposition of site-local
exciton states, sj(cid:105)'s. In other words, the extent of the
superposition represented by Eq. (3) is solely dictated by
Let us introduce X as a collective variable representing
all the degrees of freedom coupled to the exciton states.
These include the nuclear degrees of freedom of pigment
molecules as well as protein residues, and the environ-
mental electronic degrees of freedom that do not par-
ticipate directly in the electronic excitations of pigment
molecules. Then, the total Hamiltonian representing the
entirety of the exciton and its environments (the bath)
can be expressed as follows:
Hex = He + Heb(X) + Hb(X) ,
(4)
where Hb(X) is the bath Hamiltonian and Heb(X) is the
exciton-bath Hamiltonian that can be expressed as
(cid:88)
(cid:88)
(cid:88)
(cid:88)
k
j
j(cid:48)
k(cid:48)
Heb(X) =
(cid:88)
(cid:88)
=
j
k
Bjk(X)sj(cid:105)(cid:104)sk
U∗
jj(cid:48) Bj(cid:48)k(cid:48)(X)Ukk(cid:48)
ϕj(cid:105)(cid:104)ϕk . (5)
All the parameters entering the Hamiltonian Hex of Eq.
(4), δEjk's, Jjk's, and Bjk(X)'s, affect the nature of ex-
citons directly. In addition, some features of Hb(X) may
also have indirect influence on the dynamics of excitons.
In general, information on all of these is necessary for
quantitative characterization of excitons. This in turn
leads to reliable assessment of the overall exciton migra-
tion mechanism and helps elucidate the design principle
of each LHC.
However, establishing a satisfactory level of informa-
tion on all the components of Hex of Eq. (4) has remained
difficult for most LHCs. This has been true even for well
known LHCs such as the Fenna-Matthews-Olson (FMO)
complex (Fenna and Matthews, 1975) of green sulfur bac-
teria and the light harvesting 2 (LH2) complex of purple
bacteria (Cogdell et al., 2006; Hu et al., 2002), for which
the X-ray crystal structures (Fenna et al., 1977; Fenna
and Matthews, 1975; Koepke et al., 1996; McDermott
et al., 1995) were discovered more than two decades ago.
Even when main features of the exciton Hamiltonians of
these systems were understood, much still needed to be
established regarding the details of the full exciton bath
Hamiltonian. As a result, theoretical interpretations of
many spectroscopic data have remained ambiguous for a
long time.
The past decade has seen a surge of new efforts in
the spectroscopy, computational modeling, and theoreti-
cal description of delocalized excitons in LHCs. This has
become possible through advances in nonlinear electronic
spectroscopy with femtosecond time resolution, improve-
ment in computational capability, and advances in quan-
tum calculations. Another important source of motiva-
tion that has inspired a large group of recent works was
the suggestion (Collini et al., 2010; Engel et al., 2007;
Scholes, 2010) that coherent quantum dynamics of exci-
tons might play much more central role than what had
been perceived before. While definite assessment of this
suggestion remains open, especially under natural light
condition, it is true that various attempts to understand
the role of quantum coherence have led to significant ad-
vances in the spectroscopy and theoretical description of
excitons in LHCs.
3
A fundamental feature being shared by many LHCs,
which has become clearer during the past decade, thanks
to a large body of newly gathered information, is the
intermediate nature of the terms constituting Hex.
In
other words, for LHCs, it is common that δEjk's, Jjk's,
and Bjk(X)'s constituting the Hamiltonian are of com-
parable magnitudes. Typically, these parameters are on
the same order as those of the room temperature ther-
mal energy and also of the major portion of the energy
spectrum comprising Hb(X). The energetic convergence
of these multiple terms endows LHCs with a rich reper-
toire of pathways and mechanisms for exciton dynamics
and energy harvesting. At the same time, the lack of ap-
parent small parameters in these systems renders simple
perturbation theories to be rather unreliable quantita-
tively. Therefore, advanced levels of quantum dynamics
theories and computational approaches have become nec-
essary for accurate description of exciton dynamics and
relevant spectroscopic observables. In this sense, delocal-
ized excitons in LHCs have served as prominent testing
cases of modern quantum dynamics, electronic structure
calculation approaches, and spectroscopic methods.
The objective of this review is to provide a self-
contained exposition of excitons in LHCs, with particular
attention to the sources and implications of their quan-
tum delocalization. For this, we first introduce three
major LHCs for which sufficient level of experimental
information is available and extensive theoretical and
computational studies have also been made.
In order
to clarify assumptions behind prevailing models of exci-
tons in LHCs, we provide a derivation of the commonly
used form of exciton-bath Hamiltonian, and make criti-
cal assessment of underlying assumptions. Then, we offer
a comprehensive overview of major computational and
quantum dynamical methods to study and model these
excitons, and summarize applications of these methods
to the three representative systems. While this review is
devoted more to theoretical and computational aspects
as outlined above, we also provide a wide range of ex-
perimental results that are crucial for understanding the
energetics and the dynamics of excitons. We also discuss
features that are deemed important for their functional-
ity. Ultimately, this work offers both rigorous theoretical
framework and comprehensive information that can stim-
ulate future endeavor to elucidate important quantum
mechanical design principles behind efficient and robust
harvesting of excitons by LHCs.
A. FMO complex of green sulfur bacteria
4
FIG. 3 (a) FMO trimer complex. (b) View of the eight BChls
of the monomer unit A (only the heavy atoms of the BChls
ring are shown). (c) Chemical structure of the BChl pigment.
FMO complex is an LHC found in green sulfur bacteria
(Blankenship, 2014) and serves as the conduit for exci-
tons migrating to reaction centers from a large chloro-
some super-complex. Each functional unit of such a
conduit is formed by a trimer of FMO complexes as-
sembled in C3 point group symmetry, although actual
pathway of each exciton appears to be confined to one
complex for each transfer. There are 7 well-established
bacteriochlorophyll-a (abbreviated as BChl hereafter)
forming each FMO complex, which had long been be-
lieved to form a complete set of pigment molecules. How-
ever, recent findings and analyses confirmed (Schmidt am
Busch et al., 2011; Tronrud et al., 2009) the existence of
the eighth BChl as well. This serves as a linker molecule
to the chlorosome, a super-aggregate of BChls serving as
the major LHC of the green sulfur bacteria, and mediates
initial inception of the exciton into the FMO complex.
Figure 3 shows the trimer and the monomer unit of the
FMO complex and the constituting BChl.
The FMO complex has historical significance because
it is the first LHC for which X-ray crystallography struc-
tural data became available (Fenna et al., 1977; Fenna
and Matthews, 1975) and has been subject to exten-
sive spectroscopic and computational studies since then.
However, its functional significance as an LHC was con-
sidered rather minor. FMO complex does not have an
apparent internal symmetry in its arrangement of pig-
ment molecules. Thus, it was a nontrivial task to assign
complete details of its exciton states and to determine the
extent of their couplings to protein environments. Ear-
lier spectroscopic and computational investigations (Cho
et al., 2005; Louwe et al., 1997; Renger et al., 2001; Vulto
et al., 1998; Wendling et al., 2002) thus focused on quan-
tifying such parameters, as will be described in more de-
FIG. 2 A schematic of three domains constituting photosyn-
thesis, performing light harvesting (LH), charge separation
(CS), and biochemical reaction (BR). Their respective time
scales are denoted as τLH , τCS, and τBR. The length scales of
these domains are denoted as lLH , lCS, and lBR, respectively.
The ranges of energy changes that occur in these domains are
respectively denoted as ∆ELH , ∆ECS, and ∆EBR.
TABLE I Time, length, and energy scales of photosynthesis.
τLH τCS τBR lLH
100ps 1ns 1µs 100nm 10nm 1µm 0.1eV 0.5eV 1eV
lBR ∆ELH ∆ECS ∆EBR
lCS
II. OVERVIEW OF MAJOR LIGHT HARVESTING
COMPLEXES
Photosynthesis consists of three distinctive processes
as shown schematically in Fig. 2, which we call here as
light harvesting (LH), charge separation (CS), and bio-
chemical reaction (BR). Although it remains difficult to
make definite assessment of the length, time, and energy
scales of all the photosynthetic processes, it is possible
to offer rough estimates of these parameters based on
the observation of many systems studied so far, which
are listed in Table I. There is amazing diversity in how
each of these processes is executed and how the three
processes are integrated together, as has been reviewed
comprehensively down to molecular level (Blankenship,
2014). Recent efforts to model the chromatophore of
purple bacteria also offer unique insights into the syn-
ergistic organization of all the photosynthetic processes
(Cartron et al., 2014; Sener et al., 2016; Strumpfer et al.,
2012). In this review, we will focus on only one of the
three processes, namely the LH process. In this section,
we present a general introduction of three major LHCs,
Fenna-Matthews-Olson (FMO) complex of green sulfur
bacteria, light harvesting 2 (LH2) complex of purple bac-
teria, and phycobiliproteins (PBPs) of cryptophyte algae.
TABLE II Excitation energies (in cm−1) of BChls in the
for sets A1 (Wendling et al., 2002), A2
FMO complex,
(Adolphs and Renger, 2006), A3 (Schmidt am Busch et al.,
2011), T1 (Vulto et al., 1998), and T2 (Adolphs and Renger,
2006)
P. aestuari
A1
12160
A2
12230
A3
12195
C. tepidum
T1
T2
12140
12210
190
305
190
440
320
300
215
220
125
450
330
280
200
230
180
405
320
270
505
260
460
140
360
360
290
200
320
110
270
420
230
E3
E1 − E3
E2 − E3
E4 − E3
E5 − E3
E6 − E3
E7 − E3
E8 − E3
tail in the next section.
The zeroth order exciton Hamiltonian of the FMO
complex can be expressed through Eq.
(1) where the
sum runs over the 8 BChl molecules. Tables II and III
provide representative model parameters for two differ-
ent species, P. aestuari and C. tepidum, with known X-
ray crystallography structural data. The consensus es-
tablished through the earlier works was that excitons
formed in the FMO complex are modestly delocalized
(Cho et al., 2005) (up to 2-3 BChl molecules at most)
and are weakly coupled to protein environments. These
assessments were also consistent with the functional role
of the FMO complex as a "drain" of highly delocalized
excitons of the chlorosome into much more localized ones
at the reaction center.
In 2007, a 2-dimensional electronic spectroscopy
(2DES) measurement (Engel et al., 2007) reported direct
time domain observation of off-diagonal signals (in the
2DES frequency plot of signals) with beating in time that
lasts more than 500 fs at 77 K, which was suggested as be-
ing purely due to electronic coherence. This was viewed
by many researchers as surprising because direct obser-
vation of electronic coherence surviving more than 100 fs
was rare in such disordered environments. However, con-
sidering relatively small magnitudes of electronic cou-
plings (about 30 − 80 cm−1) among BChl molecules and
weak couplings of their electronic excitations to molec-
ular vibrations or surrounding protein environments, it
was not unreasonable to expect the existence of electronic
coherence with such time scale. This also appeared to be
consistent with an earlier interpretation of the oscillating
anisotropy from a pump-probe spectroscopy as being due
to electronic coherence (Savikhin et al., 1997).
However, for the electronic coherence to be detected as
real time coherent signals, dephasing due to the inhomo-
geneity should be eliminated successfully. Whether the
2DES signal observed (Engel et al., 2007) indeed rep-
resents such condition was not clear initially, let alone
the question of whether the coherent signals can indeed
5
TABLE III Electronic coupling constants of FMO complex.
The labels for sets are the same as Table II.
Jjk (cm−1)
j
1
2
3
4
5
6
7
k
2
3
4
5
6
7
8
3
4
5
6
7
8
4
5
6
7
8
5
6
7
8
6
7
8
7
8
8
P. aestuari
A1
-102
6
-6
7
-15
-14
32
8
1
14
9
-56
-2
-10
2
-69
-19
-60
89
-4
37
A2
-98.2
5.4
-5.9
7.1
-15.2
-13.5
30.5
7.9
1.4
13.1
8.5
-55.7
-1.8
-9.5
3.1
-65.7
-18.2
-58.2
88.9
-3.4
36.5
A3
-94.8
5.5
-5.9
7.1
-15.1
-12.2
39.5
29.8
7.6
1.6
13.1
5.7
7.9
-58.9
-1.2
-9.3
3.4
1.4
-64.1
-17.4
-62.3
-1.6
89.5
-4.6
4.4
35.1
-9.1
-11.1
C. tepidum
T2
T1
-87.7
-106
5.5
-5.9
6.7
-13.7
9.9
8
-5
6
-8
-4
28
6
2
13
1
-62
-1
-9
17
-70
-19
-57
40
-2
32
30.8
8.2
0.7
11.8
4.3
-53.5
-2.2
-9.6
6.0
-70.7
-17.0
-63.3
81.1
-1.3
39.7
be confirmed and validated by other independent mea-
surements (Duan et al., 2017; Thyrhaug et al., 2016).
Another key suggestion they made (Engel et al., 2007)
was that "wavelike energy motion owing to long-lived co-
herence terms" may be active for the exciton dynam-
ics rather than "semiclassical hopping mechanism," re-
viving an old debate (Fleming and Scholes, 2004; Knox,
1996). They also alluded that the FMO complex is im-
plementing a kind of quantum search algorithm. This
suggestion drew particular attention of the quantum in-
formation community and motivated exploring LHCs as
natural quantum information processing machines.
During the past decade, the FMO complex has been
subject to a new level of quantum dynamics studies, elec-
tronic structure calculations, all-atomistic simulations,
and spectroscopic measurements. These were particu-
larly focused on (i) describing the exciton dynamics as ac-
curately as possible while accounting for all major atom-
istic details of the system, (ii) rigorous understanding and
modeling of the 2DES signals and their implications, (iii)
and deeper understanding of how quantum coherence and
other quantum features make positive contribution to the
functionality of the FMO complex.
6
(a)
(b)
(c)
FIG. 4 (a) Side view of the pigment arrangement in LH2: α
and β BChl molecules of the B850 ring constitute the upper
circular aggregate whereas γ BChl molecules of B800 form the
lower circular aggregate. The carotenoids are also reported in
purple. (b) Side view of LH2 within the its membrane from a
snapshot of an all-atomistic MD simulation (the carotenoids
are not shown). (c) Geometric representation of the arrange-
ment of the BChl molecules in B800 and B850 rings adapted
with permission from (Montemayor et al., 2018). Copyright
2018 American Chemical Society. Here a reflection of 180 de-
grees has been used for clarity's sake leading the B800 ring
above the B850 ring.
B. LH2 complex of purple bacteria
LH2 complex is the primary LHC responsible for har-
vesting and delivering excitons in purple photosynthetic
bacteria. The first X-ray crystal structure (McDer-
mott et al., 1995) was determined for a bacterium called
Rhodopseudomonas (Rps.) acidophila, now reclassified as
Rhodoblastus acidophilus (Cogdell et al., 2006), reveal-
ing a cylindrical form with C9 symmetry that contains
three BChl molecules, two helical polypeptides, and a
carotenoid in each symmetry unit. Soon after, a different
LH2 complex with C8 symmetry called Rhodospirillum
(Rsp.) molischianum , now reclassified as Phaeospirillum
molischianum (Cogdell et al., 2006), was also reported
(Koepke et al., 1996). Another well-known LH2 complex
from Rhodobacter (Rb.) sphaeroides, although its struc-
ture has not been determined, is often assumed to have
almost the same structure as Rps. acidophila. All the
LH2 complexes firmly confirmed to date are known to
have CN symmetry with N = 8 − 10 only (Cleary et al.,
2013; Cogdell et al., 2006). As an LHC with the highest
level of symmetry while being finite in its size (cylin-
drical forms with both the diameter and height about
7 nm), LH2 has been subject to a wide range of spec-
troscopic studies (Cogdell et al., 2006; Sundstrom et al.,
1999). There have also been various theoretical and com-
putational studies (Hu et al., 2002) modeling/explaining
spectroscopic data and also addressing important issues
such as coherence length, dynamical time scales, and en-
ergetics of excitons.
Figure 4(a) shows the structure of the complex from
Rps. acidophila without proteins, and Fig. 4(b) depicts
the whole complex embedded in lipid membrane environ-
ments. Figure 4(c) shows a geometric representation of
LH2 (with the upside direction switched). Out of the 3N
BChl molecules constituting an LH2 with N -fold sym-
metry, N of them in the cytoplasmic side form an ex-
citon band near 800 nm wavelength and the remaining
2N in the periplasmic side form an exciton band near
850 nm wavelength at room temperature. Thus, the for-
mer is called B800 band and the latter B850 band. The
N BChl molecules constituting the B800 band are ar-
ranged circularly, with their nearest neighbor distances
about 2 nm. The 2N BChl molecules constituting the
B850 band are also arranged circularly and are formed
by two intervening concentric rings (α and β) of BChl
molecules of similar radii. In this B850 unit, the nearest
neighbor (formed by two adjacent α- and β-BChl pairs)
distances are about 1 nm.
The exciton Hamiltonian of LH2 in the absence of dis-
order can be compactly expressed as (Jang et al., 2001,
2007; Montemayor et al., 2018)
N(cid:88)
n=1
H 0
LH2 =
{Eααn(cid:105)(cid:104)αn + Eββn(cid:105)(cid:104)βn + Eγγn(cid:105)(cid:104)γn
+Jαβ(0)(αn(cid:105)(cid:104)βn + βn(cid:105)(cid:104)αn)
+Jαγ(0)(αn(cid:105)(cid:104)γn + γn(cid:105)(cid:104)αn)
+Jβγ(0)(βn(cid:105)(cid:104)γn + γn(cid:105)(cid:104)βn)}
N(cid:88)
N(cid:88)
+
(cid:88)
n=1
m(cid:54)=n
s,s(cid:48)=α,β,γ
Jss(cid:48)(n − m)sn(cid:105)(cid:104)s(cid:48)
m ,
(6)
where sn(cid:105), with s = α, β, or γ, represents the excited
state where only sn BChl is excited with its energy Es,
and Jss(cid:48)(n − m) is the electronic coupling between sn(cid:105)
and s(cid:48)
m(cid:105). The coordinates of BChl molecules and tran-
sition dipole moments can also be expressed as follows
(Jang et al., 2007; Jang and Silbey, 2003a; Montemayor
et al., 2018):
Rs,n =
Rs sin [2π(n − 1)/9 + νs]
,
Rs cos [2π(n − 1)/9 + νs]
,
sin θs cos[2π(n − 1)/9 + νs + φs]
sin θs sin[2π(n − 1)/9 + νs + φs]
Zs
cos θs
(7)
(8)
and
µs,n = µs
where s = α, β, γ and n = 1,··· , 9. Table IV provides
values of the above parameters for Rps. acidophila.
TABLE IV Geometric parameters representing the coordi-
nates and the transition dipole orientations of BChls in Rps.
acidophila. Both values derived from the crystal structure
and molecular dynamics (MD) simulation are shown (Mon-
temayor et al., 2018).
Crystal Structure MD Simulation
γ
α
β
γ
β
27.0
α
Rs (A) 25.8
Zs (A)
θs (◦)
95.8
νs (◦)
10.2
φs (◦) -108.2 64.0
94.9
-10.2
0
0
0
0
26.0 27.5 32.1
31.0
16.5
16.8
96.5 97.3 98.2
97.3
23.6
-10.2 10.2 23.3
65.2 -106.6 60.6 63.7
Table V provides data for the electronic coupling con-
stants calculated by various approaches, which will be
explained in detail in Sec. IV. For the B800 band, the
magnitudes of the nearest neighbor electronic couplings
are ∼ 20 cm−1. These electronic couplings are estimated
to be smaller, by about a factor of 2 ∼ 3, than the dis-
order in the energy and the dynamic coupling to protein
environments of the excitation of each BChl. For the
B850 band, the magnitudes of electronic couplings be-
tween nearest neighbor BChls (formed by two adjacent α
and β BChl pairs) are known to be about 200−300 cm−1.
These are comparable to the disorder in the excitation
7
energy of each BChl and are somewhat larger than the
reorganization energy of the coupling between the exci-
tation of a BChl to its environment. Therefore, excitons
in this B850 unit are more delocalized than those in the
B800 unit, although the debate on the coherence lengths
of excitons in both units still remain somewhat unset-
tled. In particular, the B850 exciton serve as a unique
example of intermediate coupling regime, where all ma-
jor energetic and dynamical parameters are of compara-
ble magnitudes. This regime defies simple classifications
normally applicable in limiting situations. For example,
there have been different estimates (Book et al., 2000;
Dahlbom et al., 2001; Jang et al., 2001; Trinkunas et al.,
2001) of the delocalization lengths of excitons ranging
from about 4 to the entire set of the BChl molecules.
TABLE V Major electronic coupling constants among BChls
in LH2 of Rps. acidophila (see Fig. 3 for notations). All
values are in cm−1. C-1: Transition dipole coupling data
based on X-ray crystal structure (Sundstrom et al., 1999).
C-2: Transition density cube method data based on X-ray
crystal structure (Krueger et al., 1998). C-3: TD-DFT/MM
data based on X-ray crystal structure (Segatta et al., 2017).
C-4: RASSCF/RASPT2 data with MM solvation based on
X-ray crystal structure (Segatta et al., 2017). C-5: TD-DFT
data with MMPol solvation based on X-ray crystal structure
(Cupellini et al., 2016). MD-1: TD-DFT data with MMPol
based on structures from MD simulation (Cupellini et al.,
2016). MD-2: TrESP calculation data based on structures
from MD simulation (Montemayor et al., 2018).
Parameter C-1 C-2 C-3 C-4 C-5 MD-1 MD-2
Jαβ(0)
Jαβ(1)
Jαα(1)
Jββ(1)
Jβα(1)
Jαβ(2)
Jγγ(1)
Jαγ(1)
Jβγ(1)
Jγα(0)
Jγβ(0)
339
317
322 238 336 563 362
288 213 288 474 409
245
140
-46 -91 -83 -87 −66 −59
-37 -63 -69 -59 −51 −29
18
14
11
20
13
13
-32 −25
42
28
27
23 −11 −11 −6 −3
3
-16 −12
-13 -19 -22 -20
−3
-10
5
-12
24
25
-22 -27 -47 -46 -50
59
23
24
22
26
44
46
-8
-9
Both the B800 and B850 units serve as active absorbing
units of photons of LH2, but the excitons formed in the
B800 unit transfer quickly to the B850 unit, which serves
as main reservoir and relayer of excitons. Pump-probe
and photon echo spectroscopy measurements (Jimenez
et al., 1996; Pullerits et al., 1997) suggested that this
transfer occurs within about 1 ps and is fairly insensi-
tive to temperature. This is much larger than theoreti-
cal estimates based on the Forster theory (Forster, 1948,
1959; Scholes, 2003), and has been the subject of theoret-
ical and experimental studies (Cheng and Silbey, 2006;
Herek et al., 2000; Jang et al., 2004, 2007; Kimura and
Kakitani, 2003; Mukai et al., 1999; Novoderezhkin et al.,
2003; Scholes and Fleming, 2000).
Other important issues concerning the LH2 complex
8
include understanding the role of carotenoids and their
dark states, and elucidating the effects of hydrogen bond-
ing on the structure and energetics of BChls. More re-
cently, the availability of the atomic force microscopy
(AFM) images of membranes containing aggregates of
LH2 complexes (Bahatyrova et al., 2004; Scheuring et al.,
2006; Scheuring and Sturgis, 2005, 2009) made it pos-
sible to investigate the nature of arrangement of LH2
complexes. Along with theoretical studies (Cleary et al.,
2013; Jang et al., 2015), this information helps address-
ing the dynamics of excitons in the aggregates of LH2
and their physical implications.
15,16-dihydrobiliverdin (DBV), whereas each β polypep-
tide chain (C or D) is linked to three phycoerythrobilins
(PEB). The corresponding pigments are labeled DBV19A,
DBV19B, PEB158C, PEB158D, PEB50/61C, PEB50/61D,
PEB82C, and PEB82D, where the subscripts denote the
protein subunit and cysteine residue linked to the chro-
mophore. The central PEB50/61D pigments are linked to
the protein by two cysteine residues. The overall PE545
structure displays a pseudo-symmetry about the 2-fold
axis relating the α1β and α2β monomers (Doust et al.,
2004).
C. Phycobiliproteins (PBP) of cryptophyte algae
Phycobiliproteins (PBPs) are primary LHCs of cryp-
tophyte algae, a diverse group of eukaryotic single-cell
photosynthetic organisms (Blankenship, 2014) that can
be found in both sea and fresh water. These organisms
live under water where spectrum of incident light lacks
the chlorophyll-absorbing region but still has significant
intensities in blue and green spectral regions. The pig-
ment molecules of PBPs are linear tetrapyrroles called
bilins, which are covalently attached to the protein scaf-
fold through single or double cysteine bonds. In the FMO
and LH2 complexes, tuning of excitation energies of pig-
ment molecules and their electronic couplings are main
factors dictating the excitonic features, respectively, as
they contain a single type of pigment.
In PBPs, how-
ever, different types of bilins can be found in the same
complex, and the possibility to tune their protonation
states or their conformation provides further mechanisms
to adapt the spectral range for efficient light-harvesting.
Another distinctive feature is that pigment molecules are
bound to proteins by covalent boding.
There are two well-known PBPs. One is phycoerythrin
545 (PE545), the structure of which was first determined
with rather low resolution (Wilk et al., 1999) and later
with higher resolution (Doust et al., 2004). The other is
called phycocyanin 645 (PC645), the detailed structure of
which was confirmed quite recently (Harrop et al., 2014).
Figure 5 shows the structures of PBPs and bilins.
The capability of PE545 to photosynthesize under low-
light conditions has been an important subject of both
experimental (Collini et al., 2010; Doust et al., 2005;
Harrop et al., 2014; Wong et al., 2012) and computa-
tional (Aghtar et al., 2017, 2014; Curutchet et al., 2011a,
2013; Hossein-Nejad et al., 2011; Huo and Coker, 2011;
Viani et al., 2014, 2013) studies. A possible explana-
tion of this efficiency is in terms of the flexible structural
nature of the pigments. This allows the optimal mod-
ulation of the absorption and energy transfer processes
through local pigment-protein interactions. The crystal
structure of PE545 contains eight bilins (Doust et al.,
2004). In particular, each α chain (A and B) contains a
FIG. 5 (a) PBP complex. (b) View of the eight bilins (only
the heavy atoms are shown) present in PE545 (black) and
PC645 (orange or grey in print). (c) Chemical structure of
the different bilins present in PE545 (PEB,DBV) and PC645
(PCB,DBV, MBV). The ellipses indicate the double bond
which is used to create the second cysteine bond to the protein
in PE545(PEB) and PC645(DBV).
Though the protein scaffolds of the PE545 and PC645
complexes are nearly identical, the compositions of pig-
ment molecules and their π-conjugation lengths are dif-
ferent. In PE545, the lowest energy states are localized
on the peripheral DVB19 molecules. For the PC645 com-
plex, which functions at lower absorption energy, the
DBV molecules reside at the dimer interface and play
the role of the source states, while the phycocyanobilin
(PCB) and the mesobiliverdin (MBV) molecules form
lower energy intermediate and sink states, respectively
(Collini et al., 2010; Huo and Coker, 2011; Lee et al.,
2017; Mirkovic et al., 2007).
We can express the exciton Hamiltonian for both
PE5454 and PC645 commonly by abbreviating the two
pigments with the highest excitation energies H1 and
H2, the two lowest ones L1 and L2, and the four in-
termediate ones as Mn, n = 1,··· , 4. Thus, for PE545,
PEB50/61C(D) and DBVs are the H and the L pairs, re-
spectively, while for PC645, DBVs are the Hs and PCBs
the Ls. With these notations, the zeroth order exciton
TABLE VI Electronic coupling constants of PE545 and
PC645. MD1: CIS/MMPol calculations on structures ob-
tained from an MD simulation (couplings are obtained as a
sum of Coulomb interactions of transition densities and an
environment term) (Curutchet et al., 2013). MD2: CIS/MM
calculations on structures obtained from an MD simulation
(couplings are obtained as dipole-dipole interactions of transi-
tion dipoles positioned at the center-of-mass of the bilin sites.
The resulting couplings were then multiplied by a value of
0.72 to take into account screening effects from the environ-
ment.) (Lee et al., 2017). C1: CIS/PCM calculations on the
crystal structure (Collini et al., 2010; Mirkovic et al., 2007).
s
s(cid:48)
H1 H2
M1
M2
M3
M4
L1
L2
H2 M1
M2
M3
M4
L1
L2
M1 M2
M3
M4
L1
L2
M2 M3
M4
L1
L2
M3 M4
L1
L2
M4 L1
L2
L1 L2
Jss(cid:48) (cm−1)
PE545
PC645
MD1
71.7
-21.5
24.5
34.0
12.1
2.2
-46.6
-15.2
19.1
-16.0
-35.6
-39.3
1.4
-6.1
7.3
6.4
-27.3
-3.7
6.8
8.2
3.5
26.3
4.0
-11.4
-36.1
34.3
11.6
-4.3
MD2
163.9
-29.5
20.6
38.8
17.9
2.7
-45.3
-22.1
29.8
-19.3
-38.1
-47.8
3.2
-6.3
7.9
6.4
-37.4
-3.5
6.2
8.7
3.6
38.7
6.2
-13.0
-48.2
45.0
13.8
-4.8
C1
319.4
-43.9
-9.6
25.3
23.8
-20.0
-46.8
7.7
43.9
29.5
30.5
21.5
48.0
4.3
86.2
3.4
53.8
-14.7
-2.9
-86.7
-15.8
49.3
7.8
29.0
-10.7
11.0
10.0
48.0
MD2
212.3
-53.4
-10.8
34.9
31.9
-19.9
-43.4
11.0
49.0
24.4
33.7
19.9
48.9
3.4
77.6
2.3
69.3
-16.0
-2.0
-78.2
-17.2
66.8
10.8
11.8
-12.4
11.2
10.7
9.8
Hamiltonian for both can be expressed as
2(cid:88)
4(cid:88)
n=1
n=1
H 0
P BP =
+
{EHnHn(cid:105)(cid:104)Hn + ELnLn(cid:105)(cid:104)Ln}
EM nMn(cid:105)(cid:104)Mn +
Js,s(cid:48)s(cid:105)(cid:104)s(cid:48) ,(9)
(cid:88)
(cid:88)
s(cid:48)(cid:54)=s
s
where s and s(cid:48) run over all possible sites. Table VI pro-
vides data for these electronic coupling constants calcu-
lated by various approaches. As LHCs with their struc-
tures known most recently among the three being con-
sidered here, spectroscopic data as well as the X-ray data
played key role early on. Detailed account of these spec-
troscopic studies are provided in the next section.
9
III. SPECTROSCOPY OF EXCITONS IN LIGHT
HARVESTING COMPLEXES
Absorption spectroscopy can identify distinctive sig-
natures of LHCs in terms of the positions and widths of
major excitonic peaks (see Fig. 1(b)). Given the struc-
tural information of an LHC, it is possible to assign some
exciton states corresponding to the peaks of the absorp-
tion lineshape directly. However, unambiguous modeling
of the entirety of the absorption line shape was difficult
in the beginning, and assignment of some peaks, espe-
cially in large LHCs without apparent symmetry, still re-
mains difficult. Excitation energies of pigment molecules
in LHCs can easily be modulated by pigment-protein in-
teractions, typically in the range of 1, 000 cm−1. De-
spite significant advances, present day electronic struc-
ture calculation methods cannot yet reliably determine
these modulations with sufficient efficiency and accuracy.
In addition, the three important factors influencing the
absorption lineshape, namely, electronic couplings be-
tween pigment molecules, vibronic couplings, and the dis-
order/fluctuations are of comparable magnitudes in the
range of 100 − 500 cm−1, making it difficult to discern
their effects on detailed features of the lineshape.
Earlier efforts to interrogate the energetics and the dy-
namics of excitons in LHCs employed conventional sub-
ensemble nonlinear spectroscopic techniques (Mukamel,
1995). For example, pump-probe and photon-echo spec-
troscopies (Cho et al., 2005; Sundstrom et al., 1999)
gave time resolved information on the response of a
subensemble of excitonic states. Hole burning spec-
troscopy (Jankowiak et al., 2011; Purchase and Volker,
2009) allowed identification of narrow zero-phonon exci-
ton states and their distributions hidden in the ensem-
ble lineshape. For LH2 and LH1 complexes of purple
bacteria and for other limited examples of LHCs, sin-
gle molecule spectroscopy (SMS) (Oellerich and Kohler,
2009; van Oijen et al., 1999; Saga et al., 2010; Schlau-
Cohen et al., 2011b) has also played an important role.
These nonlinear spectroscopic and SMS data helped ex-
tracting information on exciton relaxation kinetics, ho-
mogeneous broadening, and the disorder/fluctuations.
However, they in general fell short of offering unambigu-
ous interpretation of experimental signals for excitons in
LHCs, the complexity of which typically causes multi-
tudes of competing models/scenarios to be viable for in-
terpreting the spectroscopic signals.
Further advances in laser technology have made it pos-
sible to conduct a general 2DES (Cho, 2008; Fuller and
Ogilvie, 2015; Jonas, 2003) and allowed investigating new
energetic and temporal details of exciton dynamics in
LHCs (Ginsberg et al., 2009; Schlau-Cohen et al., 2011a).
2DES is a femtosecond pump-probe technique that re-
solves both pump and probe frequencies. It thus corre-
lates the visible-light absorption spectrum, which enables
related excitonic states to be identified by cross-peaks.
2DES is similar to transient absorption spectroscopy.
However, two excitation pulses are used cooperatively
to excite the sample followed by a third "probe-pulse"
which interacts with the sample after the pump-probe
time delay. More detailed account of these spectroscopic
studies are outlined below for the FMO complex, LH2
complex, and the PBPs of cryptophyte algae.
1. Spectroscopy on FMO complex
FIG. 6 Experimental and theoretical absorption spectra of
FMO complex at three different temperatures adapted from
(Renger et al., 2012) with permission. Copyright 2012 Ameri-
can Chemical Society. The two simulation data in the middle
and right panel show that improvement in the bath spectral
density brings qualitative features of theoretical line shapes
closer to those of experimental ones.
Already in 1990s, high quality steady state linear spec-
troscopic data for FMO complexes such as absorption,
linear dichroism (LD), and circular dichroism (CD) be-
came available (Louwe et al., 1997; Miller et al., 1994;
Savikhin and Struve, 1996; Vulto et al., 1998; Wendling
et al., 2002). Earlier efforts (Gulen, 1996; Louwe et al.,
1997; Pearlstein, 1992; Vulto et al., 1998; Wendling et al.,
2002) to explain these spectroscopic data were based on
simple exciton models of the form of Eq. (1), without
including bath interactions. Thus, line broadening due
to environmental relaxation of exciton states and inter-
exciton dynamics were not properly taken into consider-
ation in these works. Nonetheless, the fittings of spectral
data and the resulting model parameters turned out to
be reasonable. This indicates that the exciton-bath cou-
pling and inter-exciton couplings in the FMO complex
are not dominant factors. Later, the quality of fitting
was improved by including proper lineshape functions
and utilizing more systematic fitting algorithms (Adolphs
and Renger, 2006). The model parameters obtained from
these fittings are compared in Tables II and III. Figure
10
6 shows experimental absorption lineshapes of the FMO
complex compared with a recent set of theoretical line-
shapes (Renger et al., 2012). More detailed account of
computational methods underlying these theoretical cal-
culations will be provided in the later sections of this
work.
Subensemble nonlinear spectroscopies such as pump
probe (van Amerongen and Struve, 1991; Buck et al.,
1997; Freiberg et al., 1997; Savikhin et al., 1997; Savikhin
and Struve, 1994, 1996), hole burning (Johnson and
Small, 1991; Ratsep et al., 1999), and photon echo
(Prokhorenko et al., 2002) were employed to gain infor-
mation on the dynamics and the energetics of exciton
states that are not clearly visible in ensemble linear spec-
troscopic data. These experiments revealed signatures of
relaxation dynamics ranging from sub-hundred femtosec-
onds to hundred picoseconds. However, definite assess-
ment and quantitative quantum dynamical modeling of
these data have not been pursued extensively.
Advances in 2DES technique made it possible to ac-
cess new information on exciton states of the FMO com-
plex that had not been available otherwise. Earlier suc-
cess was made in determining detailed excitonic path-
ways (Brixner et al., 2005; Cho et al., 2005) that were
largely consistent with a model exciton Hamiltonian de-
veloped earlier (Vulto et al., 1998, 1999). This was soon
followed by direct observation of beating signals lasting
more than 500 fs at 77 K (Engel et al., 2007). Engel and
coworkers made further progress, and have reported more
detailed experimental results (Hayes and Engel, 2011;
Hayes et al., 2010) including the evidence that beating
signals can be observed even at room temperature (Pan-
itchayangkoon et al., 2010). While these were exciting
results that have motivated a broad community of experi-
mentalists and theoreticians, interpretation of such 2DES
signals was not straightforward because of the fact that
the FMO complex has more than two exciton states and
the exciton-bath couplings are more complicated than
simple models that are typically used for analyzing 2DES
signals. There was strong possibility that the beating
signals might have originated from any of electronic, vi-
brational, and vibronic contributions. In fact, whether
all of these contributions can be discerned by any 2DES
was not clear at all (Butkus et al., 2012). Thus, there
has been ongoing debate and various analyses on the ori-
gins and sources of the beating signal (Christensson et al.,
2012; Fransted et al., 2012; Fujihashi et al., 2015; Liu and
Kuhn, 2016; Plenio et al., 2013; Tempelaar et al., 2014;
Tiwari et al., 2013). Furthermore, such beating signals
have not yet been independently confirmed in different
2DES data by other groups (Duan et al., 2017; Thyrhaug
et al., 2016).
Due to the unique arrangement of pigment molecules
in the FMO complex, different exciton states have well-
defined directions of transition dipole moments. Thus,
control and information of the light polarization can play
an essential role in characterizing the dynamics directly
associated with electronic coherence. Indeed, utilization
of the information on polarizations of light pulses has
been shown to be critical in deducing structural data
(Read et al., 2008). More recent work demonstrated that
full polarization control allows accurate description of
electronic structure and population dynamics (Thyrhaug
et al., 2016).
With further advances in spectroscopic and sample
preparation techniques, efforts to clarify detailed char-
acteristics of excitons have continued in different direc-
tions. A new hole burning spectroscopy of FMO trimer
and modeling in terms of excitonic calculations were re-
ported (Kell et al., 2014). Transient absorption and time-
resolved fluorescence spectroscopies were conducted for
several mutants of C. tepidum (Magdaong et al., 2017).
A new 2DES spectroscopy provided evidence that FMO
indeed functions as a conduit of exciton energy (Dost´al
et al., 2016). A single molecule spectroscopy has finally
become possible for the FMO complex (Lohner et al.,
2016).
2. Spectroscopy on LH2
Thanks to the high symmetry of the LH2 complex, the
number of independent parameters minimally necessary
for fitting its lineshape is relatively small compared to
the total number of pigment molecules involved. Indeed,
the average values of excitation energies of α-, β-, and
γ-BChls, nearest neighbor electronic couplings, and the
magnitudes of the disorder for each of the B800 and B850
units seem sufficient for fairly accurate fitting of ensemble
lineshape. However, on the other hand, the easiness of
fitting the ensemble lineshape has also been a contribut-
ing factor for the lack of clear consensus on some details
of the exciton Hamiltonian.
For example, modeling of both absorption and CD
spectra (Georgakopoulou et al., 2002) based on the crys-
tal structure support the assumption that the excitation
energy of α-BChl is about 300 cm−1 higher than that
of β-BChl. However, more recent computational studies
(Cupellini et al., 2016; Montemayor et al., 2018) suggest
that the conformations and local environments of α- and
β-BChls are virtually the same in membrane environ-
ments without indicating significant difference in their
excitation energies. Efforts to refine the exciton-bath
model of LH2 by combining more comprehensive spec-
troscopic data and advanced computational tools have
continued, for example, based on temperature depen-
dent absorption spectra (Urboniene et al., 2007; Zer-
lauskiene et al., 2008) and combination of temperature
dependent absorption, fluorescence, and fluorescence-
anisotropy (Pajusalu et al., 2011a). As yet, there has
not been any attempt to explain all existing steady state
spectroscopic data over all the temperatures based on a
11
universal exciton model.
The determination of the X-ray crystal structures of
LH2 complexes coincided with new advances in nonlinear
spectroscopic theories and techniques. Thus, LH2 soon
became an important testing ground for evolving nonlin-
ear spectroscopic techniques. Along with the standard
pump-probe and hole burning spectroscopies, various
versions of four wave mixing optical spectroscopy have
been used to interrogate mechanistic details and time
scales of exciton dynamics (Lampoura et al., 2000; Sund-
strom et al., 1999). Ultrafast fluorescence up-conversion
(Jimenez et al., 1996) provided information on time scales
of exciton dynamics within B800 and B850 units. It also
showed that the time scale of the exciton transfer from
the B800 unit to B850 at room temperature is about
1.5 ps. Pump-probe spectroscopies at a few different
temperatures (Pullerits et al., 1997) also provided sim-
ilar estimates, and demonstrated that the transfer time
decreases with temperature moderately from 1.5 ps at 4.2
K to 0.7 ps at 300 K. Time resolved transient absorption
spectroscopy of isolated and native membrane-embedded
LH2 complexes of Rb. sphaeroides at 10 K (Timpmann
et al., 2000b) showed that the inter-LH2 exciton transfer
time is larger than 1 ps. Three pulse photon echo peak
shift suggested that (Agarwal et al., 2002) the intra-LH2
exciton relaxation takes about 200 fs, and the inter-LH2
exciton transfer about 5 ps. Combination of various spec-
troscopic techniques were made to interrogate the delo-
calization length of excitons and its dynamical localiza-
tion with time scales less than 200 fs (Book et al., 2000).
As in the case of the FMO complex, real time coher-
ent beating signal was observed in 2DES data of LH2
complex. Angle resolved coherent four wave mixing spec-
troscopy showed evidence for coherent dynamics that can
be well separated from the relaxation signal due to en-
ergy transfer (Mercer et al., 2009). 2DES has also been
used to determine parameters of the Hamiltonian for LH3
complex (Zigmantas et al., 2006), which is similar to LH2
and appears under low light condition. Clear peaks cor-
responding to exciton transfer from the B800 unit to the
B850 unit were shown to emerge even at about 200 fs
after excitation (Harel and Engel, 2012). Recent 2DES
spectroscopy suggested new aspects on the possible role
of dark states of carotenoids (Ostroumov et al., 2013)
and dark charge transfer states of BChls (Ferretti et al.,
2016) in energy transfer dynamics.
Hole burning spectroscopy (Jankowiak et al., 2011;
Purchase and Volker, 2009; Wu et al., 1997) has long
played a unique role in gaining quantitative information
on the low lying exciton states, and has helped validating
the exciton Hamiltonian model for LH2. However, more
direct experimental demonstration of the validity of the
Frenkel exciton model for LH2 came from low tempera-
ture single molecule spectroscopy based on fluorescence-
excitation technique (Berlin et al., 2007; Brotosudamo
et al., 2009; Kunz et al., 2012, 2014; Lohner et al., 2015;
12
et al., 2011).
3. Spectroscopy on PBPs of cryptophyte algae
FIG. 7 Single LH2 fluorescence-excitation spectra (solid black
lines) and emission spectra (dashed blue and red lines, or
dashed grey lines in print) provided by Jurgen Kohler. The
corresponding ensemble lineshapes are shown in the top. The
left panel
is for LH2 complexes with sharp emission line
shapes, and the right panel is for LH2 complexes with broad
emission line shapes. Detailed experimental method and more
comprehensive set of data are available in (Kunz et al., 2012).
van Oijen et al., 1999). In particular, the SMS lineshapes
for the B850 unit clearly consisted of two major exciton
peaks (k = ±1), which are perturbed due to the disorder,
and one or two minor exciton peaks at higher energies.
Figure 7 shows line shapes reported from a recent SMS
experiment (Kunz et al., 2012).
Following the success of low temperature SMS exper-
iments, room temperature SMS experiments were also
conducted to investigate the nature and the dynamics
of excitons at physiological temperature (Bopp et al.,
1997, 1999). However, the noisiness of the data made it
difficult to obtain any definite microscopic information.
More recently, room temperature single molecule emis-
sion spectroscopy of LH2 was shown to be able to iden-
tify three emissive states switching at room temperature
(Schlau-Cohen et al., 2011b). Single molecule femtosec-
ond pump probe spectroscopy employing ultrafast phase
coherent excitation was also reported recently (Hildner
et al., 2013).
How excitons migrate in aggregates of LH2 complexes
has important implication for understanding the design
principle of efficient energy conversion. To this end, de-
tailed information on the arrangement of LH2 complexes
is needed. AFM images have revealed various patterns
of aggregates depending on light conditions (Scheuring
and Sturgis, 2005). The extents of order and disorder
also vary with other growth conditions and species (Olsen
et al., 2008; Sturgis et al., 2009). AFM images and pump
probe spectroscopy of aggregates reconstituted in phos-
pholipids have become available (Sumino et al., 2013).
Time resolved spectroscopy of exciton migration in ar-
rays of LH2 complexes has also been reported (Pflock
FIG. 8 Spectroscopic data for PC645 adapted from (Dean
et al., 2016) with permission from Elsevier. (a) Absorption
lineshapes at 77 K (dashed black line) and ambient temper-
ature (solid black line), and fluorescence lineshape (solid red
line, or grey line in print) at ambient temperature. (b) Vi-
brational frequencies extracted from transient absorption by
Fourier transform at three different emission frequencies. (c)
2DES taken at 295 K. t2 refers to population time. (d) 2DES
taken at 77 K.
Linear ensemble spectroscopy, polarization anisotropy,
and transient grating were employed early on as a tool
to augment X-ray crystallography to construct an exci-
ton model for PE545 (Doust et al., 2004). Oscillations
lasting up to about 1 ps were observed, which were as-
signed mostly coming from vibrational coherence (Doust
et al., 2004). Transient absorption spectroscopy (Doust
et al., 2005) also revealed broad time scales of energy
transfer within the PE545 ranging from 250 fs to picosec-
onds. A comprehensive set of absorption, CD, fluores-
cence, and time resolved transient absorption at both
77 K and 300 K were reported (Novoderezhkin et al.,
2010). This work also employed the experimental results
to construct the exciton Hamiltonian and model exciton
transfer kinetics.
Photon echo spectroscopy (Collini et al., 2010) of both
PE545 and PC645 at 294 K reported oscillatory cross
peaks in the two dimensional representation lasting more
than 400 fs, which were initially interpreted as originating
from electronic coherence. Further interrogation of the
sources and implications of these signals have continued,
and ensuing studies (McClure et al., 2014; Turner et al.,
2012) clarified that most of them have vibrational origin,
confirming the earlier suggestion (Doust et al., 2004). In
fact, this was not surprising considering the covalent na-
ture of the pigment-protein bonding and rather strong vi-
bronic coupling (Kolli et al., 2012). Recent studies (Dean
et al., 2016; O'Reilly and Olya-Castro, 2014) suggest that
vibronic couplings are in fact being utilized positively for
efficient and robust light harvesting capability. Figure
8 provides the absorption and fluorescence lineshapes of
PC645, vibrational spectra obtained from different sec-
tions of transient absorption, and 2DES profiles at two
different temperatures.
IV. EXCITON-BATH HAMILTONIAN
The fact that the Frenkel exciton Hamiltonian of Eq.
(1) provides reasonable description of the electronic spec-
tra of many LHCs was well established at phenomeno-
logical level. However, computational efforts to vali-
date its assumption and to determine the parameters di-
rectly through first principles calculation are fairly recent
and are relatively in early stage. For further progress
in this direction, it is important to clarify assumptions
and approximations involved in the exciton Hamiltonian,
or more generally, the exciton-bath Hamiltonian (EBH).
To this end, we here provide a comprehensive review of
quantum mechanical assumptions implicit in the EBH
typically used for LHCs and computational methods to
calculate key elements of the EBH such as the excitation
energies, electronic couplings, and the spectral densities
of the bath.
A. Derivation
1. Aggregates of pigment molecules
Consider an aggregate of pigment molecules or chro-
mophores (pigment molecules or part of them) more gen-
erally. The total molecular Hamiltonian representing the
aggregate can in general be expressed as
Nc(cid:88)
j=1
Hc =
Hj +
1
2
Nc(cid:88)
Nc(cid:88)
j=1
k(cid:54)=j
Hjk ,
(10)
where Nc is the total number of chromophores, Hj is the
full molecular Hamiltonian of the jth chromophore, and
Hjk is the interaction Hamiltonian between the jth and
the kth chromophores.
For the jth chromophore, we denote the positions and
momenta of the ith electron as rj,i and pj,i, and those
of the lth nucleus as Rj,l and Pj,l. Given that the jth
chromophore has Lj nuclei and Nj electrons, Hj in the
13
Zj,lZj,l(cid:48)e2
Rj,l − Rj,l(cid:48)
above equation can be expressed as
P2
j,l
2Mj,l
+
l=1
Lj(cid:88)
Nj(cid:88)
Nj(cid:88)
i=1
1
2
−
p2
j,i
2me
Nj(cid:88)
i=1
i(cid:48)(cid:54)=i
Hj =
+
+
Lj(cid:88)
l(cid:48)(cid:54)=j
Lj(cid:88)
Nj(cid:88)
Lj(cid:88)
1
2
l=1
l=1
i=1
e2
rj,i − rj,i(cid:48)
Zj,le2
Rj,l − rj ,i
= Tj,n + Vj,nn + Tj,e + Vj,en + Vj,ee ,
(11)
where the terms following the second equality are ab-
breviations of the kinetic ( T ) and potential ( V ) energy
operator terms. The subscripts n and e in these terms
respectively refer to nuclear and electron degrees of free-
dom.
The second term in Eq. (10) represents interactions
between all chromophores. These are pairwise at the
level of explicit description of all Coulomb interactions
among electrons and nuclei. Namely, each component,
Hjk, represents the interaction between the jth and the
kth chromophores, and consists of four potential terms
as follows:
l=1
Lj(cid:88)
Lj(cid:88)
Nj(cid:88)
l=1
l(cid:48)=1
Lk(cid:88)
Nk(cid:88)
Nk(cid:88)
i=1
i=1
i(cid:48)=1
Hjk =
−
+
Zj,lZk,l(cid:48)e2
Rj,l − Rk,l(cid:48)
Rj,l − rk,i − Lk(cid:88)
Zj,le2
l=1
Nj(cid:88)
i=1
Zk,le2
Rk,l − rj,i
e2
rj,i − rk,i(cid:48)
= Vjk,nn + Vjk,ne + Vjk,en + Vjk,ee ,
(12)
where each term in the last line is again an abbreviation
of the corresponding interaction potential term.
One can employ the standard quantum mechanical
procedure to derive an EBH corresponding to Eq. (11).
Appendix A provides a detailed description of such pro-
cedure, based on the assumption that the ground elec-
tronic state and the site excitation state can be defined
in terms of direct products of adiabatic electronic states
of independent chromophores defined at reference nuclear
coordinates. The default choice for these reference nu-
clear coordinates are those of the optimized ground elec-
tronic states of chromophores. Thus, the ground elec-
tronic state and the site excitation states are defined by
Eq. (A29) and Eq. (A30). Collecting all the terms in
Appendix A, Hc can thus be expressed as
Nc(cid:88)
j=1
Hc = Ec,0
g g(cid:105)(cid:104)g +
Nc(cid:88)
(cid:88)
Ec,0
j=1
j
+
+
sj(cid:105)(cid:104)sj +
gj g(cid:105)(cid:104)sj)
jg sj(cid:105)(cid:104)g + J c,0
( J c,0
(cid:88)
(cid:88)
jk sj(cid:105)(cid:104)sk
J c,0
k(cid:54)=j
j
Bc,0
j
sj(cid:105)(cid:104)sj + H c,0
b
.
(13)
j
g
In the above expression, Ec,0
is the energy of the ground
electronic state and is defined by Eq. (A36). J c,0
jg and
J c,0
gj are the electronic couplings between the site excita-
tion state sj(cid:105) and the ground electronic state g(cid:105), and are
defined by Eqs. (A33) and (A34). These terms originate
from interactions between electrons in the excited state of
the jth chromophore and the nuclear degrees of freedom
in the ground electronic state of all others. Note that
these are in general operators with respect to the nuclear
degrees of freedom and may be responsible for nonadia-
batic transitions, although small. Ec,0
is the electronic
energy of the site excitation state sj(cid:105) as defined by Eq.
(A37), J c,0
jk is the electronic coupling between site exci-
tation states sj(cid:105) and sk(cid:105), and is defined by Eq. (A35).
represents the coupling between sj(cid:105) and all the nu-
Bc,0
clear degrees of freedom, as described by Eq. (A39), and
H c,0
represents the bath Hamiltonian originating from
the nuclear degrees of freedom of all the chromophores.
The definition of this is given by Eq. (A38).
b
j
j
It is important to note that we have labeled all the
terms defined in Eq. (13) with an additional superscript
0. This was to make it clear that they are defined with re-
spect reference adiabatic states of independent pigment
molecules.
In fact, even for simple aggregates of pig-
ments, there are many-body effects of other pigments,
which affect the definition and calculation of adiabatic
electronic states of each pigment molecule, making the
actual values of parameters and Hamiltonian terms dif-
ferent from the zeroth order ones in Eq. (13). For LHCs,
these implicit effects due to other pigment molecules are
expected to be less significant than those due to pro-
tein environments in general, which can be considered
together.
2. Light harvesting complex
The presence of the protein environment in the LHC
affects the molecular Hamiltonian of the aggregate in all
of its terms, and adds an additional one referring to its
own degrees of freedom. Expressing all these in the site
excitation basis, one can obtain the following expression
for the Hamiltonian of the LHC:
14
gj)g(cid:105)(cid:104)sj(cid:111)
gj + J r
g )g(cid:105)(cid:104)g
jg)sj(cid:105)(cid:104)g + ( J c
(cid:88)
(cid:88)
k(cid:54)=j
j
HLHC = (Ec
g + Er
( J c
jg + J r
(cid:110)
j=1
Nc(cid:88)
Nc(cid:88)
Nc(cid:88)
j=1
+
+
+
Nc(cid:88)
(Ec
j + Er
j )sj(cid:105)(cid:104)sj +
(J c
jk + J r
jk)sj(cid:105)(cid:104)sk
( Bc
j δjk + Br
jk)sj(cid:105)(cid:104)sk + H c
b + H r
b ,
(14)
j=1
k=1
where δjk in the last line is the Kronecker-delta symbol.
All the terms with superscript c represents contributions
from chromophores, which also include all the implicit
effects on the adiabatic state of each chromophore by the
protein environments and the other chromophores in the
ground electronic state. The terms with superscript r
represent explicit contributions of the protein environ-
ment.
Equation (14) is the final and the most general form of
the exciton-bath Hamiltonian for LHC, and includes all
possible interactions between single excitons and other
degrees of freedom. The terms that are missing here are
intra-pigment non-adiabatic terms, spontaneous emission
terms of the site excitation states and interaction terms
with the radiation, which will be considered later. The
common assumption implicit in most theoretical models
developed so far is that the environment induced nonadi-
abatic coupling between the ground and the exciton state
(cross terms between g(cid:105) and sj(cid:105) of Eq. (14)) are neg-
ligible compared to the spontaneous emission term and
the interaction terms with the radiation. Thus, we also
assume that they are negligible.
Combining the contributions of the chromophores and
protein environments, we define the following parame-
ters:
Eg = Ec
Ej = Ec
Jjk = J c
Bjk = Bc
Hb = H c
g + Er
g
j + Er
j ,
jk + (cid:104) J r
jk(cid:105) ,
j δjk + Br
jk + J r
b + H r
b .
jk − (cid:104) J r
jk(cid:105) ,
(15)
(16)
(17)
(18)
(19)
Employing the above definitions and neglecting the envi-
ronment induced nonadiabatic terms between the ground
electronic and the single exciton states, as noted above,
Eq. (14) can be simplified as
HLHC = Egg(cid:105)(cid:104)g + He +
= Egg(cid:105)(cid:104)g + Hex ,
Bjksj(cid:105)(cid:104)sk + Hb
j=1
k=1
(20)
where He has the same form as Eq. (1) and the second
equality serves as the definition of Hex. As can be seen
Nc(cid:88)
Nc(cid:88)
above, this also has the same form as Eq. (4) except that
the variables X is not explicitly shown. The full determi-
nation of HLHC given by Eq. (20) is a challenging task in
general. Furthermore, even if the full information on the
Hamiltonian of the above type were available, the actual
quantum dynamical calculation poses as another theo-
retical challenge. Thus, in consideration of practicality,
further approximations have been made.
3. Site diagonal coupling to bath of harmonic oscillators
Given the structural and spectroscopic stability of
LHCs, it is reasonable to assume that the displacements
of the bath degrees of freedom remain small enough to
be governed by almost linear restoring forces. Under this
condition, the bath Hamiltonian can be approximated as
a set of harmonic oscillators, and the exciton-bath inter-
action terms can be assumed to be linear in the bath co-
ordinates. While these interaction terms may in general
involve both diagonal and off-diagonal terms of the exci-
ton Hamiltonian in the site excitation basis, most models
of LHCs developed and considered so far have assumed
the existence of only diagonal couplings. At least for the
case of the FMO complex, explicit calculation of normal
modes confirmed that (Renger et al., 2012) off-diagonal
exciton-bath couplings are about an order of magnitudes
smaller than diagonal ones, offering microscopic justifi-
cation for such assumption.
Thus, the typical form of the EBH that has been used
for LHC so far assumes that Bjk is diagonal and linear
in the displacement of bath coordinates. Within this ap-
proximation, Eq. (20) can be expressed as
ωngj,n(bn + b†
(cid:19)
b†
n
bn +
1
2
,
(21)
where b†
n and bn are raising and lowering operators of
the nth normal mode constituting all the bath degrees
of freedom. This approximation also ignores Duschin-
sky rotation, which can be significant if excitation causes
nontrivial structural change of chromophores.
Even though the exciton-bath coupling is assumed to
be diagonal in the site excitation basis, certain delocal-
ized vibrational normal modes can be coupled to different
site excitation states together, which are called common
modes. Thus, in general, the complete specification of
the exciton-bath coupling requires specification of the fol-
lowing spectral densities of the bath for all pairs of j and
Ejsj(cid:105)(cid:104)sj +
Jjksj(cid:105)(cid:104)sk
Nc(cid:88)
(cid:88)
j=1
k(cid:54)=j
n)sj(cid:105)(cid:104)sj
Nc(cid:88)
j=1
HLHC ≈ Egg(cid:105)(cid:104)g +
(cid:88)
Nc(cid:88)
(cid:88)
j=1
n
+
+
(cid:18)
n
ωn
15
k:
Jjk(ω) = π(cid:88)
n
δ(ω − ωn)ω2
ngj,ngj,k .
(22)
For practical calculations, it is convenient to decompose
the spectral density into the sum of contributions from
the intramolecular vibrational modes of the given chro-
mophore itself and those of the environments. It is rea-
sonable to assume that the former is local to each site
excitation state in general. Thus, Eq. (22) can be de-
composed into
Jjk(ω) = J c
j (ω)δjk + J r
jk(ω) .
(23)
The reorganization energy of the bath upon the creation
of each site excitation is thus given by
λj = λc
j + λr
j ,
where
(cid:90) ∞
(cid:90) ∞
0
0
λc
j =
λr
j =
1
π
1
π
J c
j (ω)
ω
J r
j (ω)
ω
,
.
dω
dω
B. Electronic structure calculation
1. Excitation energies
(24)
(25)
(26)
The EBH for LHC given by Eq. (20) or (21) serves as
an efficient framework for describing excitons in LHCs,
and can be used for practical calculations even for large
aggregates of many interacting pigments by keeping the
computational cost limited (Curutchet and Mennucci,
2017). Moreover, as the cost is mainly due to the calcu-
lation of localized excitations, accurate quantum chemi-
cal methods can in principle be used within this model.
However, even with this simplification, typically large
molecular dimensions of the pigments present in natural
LHCs (around 50-60 heavy atoms) prevent routine use of
highly accurate methods such as ab initio multi-reference
or coupled-cluster methods.
A good compromise between cost and accuracy is rep-
resented by the Density Functional Theory (DFT) and
its excited-state extension, commonly known as Time-
Dependent (TD) DFT. Beside its computational effi-
ciency, this approach has two main limitations, (i) to be
highly sensitive to the selected functional and (ii) to be
a single-reference description. The first issue is particu-
larly relevant for excitations that have a (partial) charge-
transfer character. TD-DFT is in fact well known to be
inaccurate in those cases due to its intrinsic limit, com-
monly known as "self-interaction error"; this arises from
the spurious interaction of an electron with itself in the
Coulomb term of the DFT Hamiltonian, which is not ex-
actly canceled by the exchange contribution, e.g., as in
the Hartree-Fock (HF) approach.
More recently, functionals which introduce range sep-
aration into the exchange component and replace the
long-range portion of the approximate exchange by the
HF counterpart, have been proposed to correct this error
(Tsuneda and Hirao, 2014). By using these long-range
corrected functionals, CT-like transitions can be semi-
quantitatively described at TD-DFT level. The second
issue, namely of being a single-reference approach, is in-
stead particularly relevant for excitations in long conju-
gated systems such as carotenoids. In those systems, in
fact, a multi-reference approach is compulsory if a cor-
rect evaluation of the different ("dark" and "bright") ex-
cited states is needed. For such systems, good perfor-
mances have been shown by the multi-reference exten-
sion of DFT, known as DFT/MRCI approach (Andreussi
et al., 2015; Grimme and Waletzke, 1999; Kleinschmidt
et al., 2009).
The optimal electronic structure calculation method
for LHCs should not only give accurate excitation ener-
gies for the different pigments, but should also provide
correct description of the variation of the electronic den-
sity upon excitation and the transition density. The for-
mer is required for accurate evaluation of excited state
properties such as the geometrical gradients, which are
used for calculating relaxed geometries and the coupling
between electronic and nuclear degrees of freedom (see
below). The latter (transition density) determines not
only the transition dipole corresponding to a selected
excitation but also the inter-pigment interactions defin-
ing the electronic coupling, Jjk (see below). These re-
quirements on the electronic densities largely limit the
range of methods that can be used safely. For exam-
ple, many semi-empirical approaches, which have been
largely used in the past, should be used with care.
In
fact, by construction (e.g. by parameterization), they
often give correct excitation energies but this does not
mean that the corresponding transition densities are cor-
rect as well. Moreover, only by chance, they could give a
reliable description of the variation of the electronic den-
sity upon excitation, as excited state properties are not
used in their parameterization. Once more, at present,
TD-DFT represents the only feasible approach available
in most cases even if the same two limitations, (i) and
(ii), mentioned in regard to the excitation energies also
apply for excited state densities and transition densities.
As a matter of fact, some DFT functionals can give
quite accurate transition energies but may completely fail
in correctly describing the change in the electronic den-
sity upon excitation. In general, transition densities are
known to be less sensitive to errors of the DFT functional
than excited state densities (Munoz-Losa et al., 2008).
However, benchmark data should always be used to be
fully confident about the selected functional. Another as-
16
pect that should be preliminarily checked before selecting
an electronic structure calculation method is the robust-
ness with respect to the change of geometry. Some calcu-
lation methods, especially those based on semi-empirical
formulation, can behave very differently if the molecular
system is out of the minimum region of the potential en-
ergy surface. For pigment molecules in LHCs, significant
deviations from the minimum geometry are possible due
to the temperature-dependent effects and/or to geomet-
rical constraints due to the protein matrix. Indeed, this
is a very important aspect to consider when selecting a
reliable electronic structure calculation method. In this
respect, at present, systematic studies providing satis-
factory references are not available to the extent of our
knowledge.
2. Electronic couplings
The electronic coupling Jjk in Eq. (1), (20), or (21) is
a key quantity determining both the dynamics and the
mechanism of exciton transfer. The magnitude of the
electronic coupling, compared to the exciton-phonon in-
teractions, determines whether states localized on differ-
ent sites are mixed to generate delocalized exciton states.
An accurate computation of the electronic coupling is
thus necessary not only to predict the rates of energy
transfer processes, but also to determine their mecha-
nism. As shown in Eq. (17), Jjk consists of two terms,
one direct interaction term between chromophores, J c
jk
jk(cid:105). Here-
and the other environmental contribution, (cid:104) J r
after, we denote this latter contribution simply as J r
kj.
Thus, Eq. (17) can be expressed as
Jjk = J c
jk + J r
jk .
(27)
Within the derivation of Appendix A, the direct elec-
tronic coupling between chromophores J c
jk is defined by
generalizations of Eqs.
(A35) and (A28) that include
the implicit effect of environments as well. Following the
convention, this can also be expressed as
(cid:90)
J c
jk =
(28)
drdr(cid:48)ρtr∗
j
(r)ρtr
e2
k (r(cid:48))
r − r(cid:48) ,
k (r(cid:48)) = ρ10
j
(r) = ρ01
j (r) and ρtr
k (r(cid:48)), transi-
where ρtr∗
tion densities of chromophores j and k including the
implicit effects of environments. Thus, this represents
the Coulomb coupling between de-excitation of the jth
chromophore and the excitation of the kth chromophore
within the given LHC. More generally, additional terms
due to exchange interactions and overlap between or-
bitals of different chromophores should be included. This
is missing in Eq.
(28) because it has been assumed
that electrons from different chromophores are distin-
guishable and have zero overlap. For most LHCs, the
inter-chromophore distances are far enough to make the
Coulomb term by far the most dominant one. The ap-
proximation of Eq. (28) is well justified under such con-
dition. Of course, short-range non-Coulomb interactions
are not always negligible. For example, contributions
originating from charge-transfer states were shown to
make modest contribution to the nearest-neighbor cou-
plings in the B850 unit of LH2 (Scholes et al., 1999) and
to constitute significant portion of the interaction be-
tween the special pairs of purple bacteria and PS1 (Mad-
jet et al., 2009), where the BChls are closely packed to-
gether.
The simplest but still widely used approximation for
the Coulomb coupling is to use the expansion of the tran-
sition densities up to dipolar terms, which leads to the
following point transition dipole approximation:
J c,dp
jk =
µj · µk
r3
jk
− 3(µj · rjk)(µk · rjk)
r5
jk
,
(29)
where rjk is the distance vector between pigments j and
k, and µj and µk are the corresponding electronic tran-
sition dipole moments.
The dipolar term given by Eq. (29) yields the well
known r−3
jk asymptotic dependence of the singlet elec-
tronic coupling. Where applicable, the dipole approxi-
mation has the clear advantage of only needing experi-
mental data, namely the transition dipole moments and
the distance between the centers of pigment molecules,
provided that the orientation of the transition dipole mo-
ments is known. For this reason, the dipole approxi-
mation has been widely employed (Adolphs and Renger,
2006; Damjanovi´c et al., 2002; Georgakopoulou et al.,
2002; Jang et al., 2001; Wan et al., 2014). However, par-
ticular care should be taken in using the approximation
because it can result in significant error in the estima-
tion of the Coulomb coupling between proximate pigment
molecules. For example, this is the case for the electronic
couplings between the nearest neighbor BChl molecules
in the B850 unit of LH2 complex and for the two central
bilins of PBPs of cryptophyte algae. On the other hand,
for FMO complex, the transition dipole approximation
serves as reasonable approximation for all the electronic
couplings.
Another widely used method to compute J c
jk is based
on the projection of the transition densities onto atomic
transition charges (TrCh). The Coulomb coupling is then
computed as the electrostatic interaction between those
charges:
(cid:88)
n∈j
m∈k
J c,TrCh
jk
=
qn qm
rnm
,
(30)
where the indices n and m run over the atoms of j and
k, respectively, qn and qm are the transition charges of
atoms n and m, and rnm is the distance between them.
Atomic transition charges with various definitions have
17
been used for a long time to compute Coulomb couplings
(Carbonera et al., 1999; Chang, 1977; Duffy et al., 2013).
Arguably, a definition of atomic charges that is physi-
cally accurate and adequate for electrostatic interaction
is the one based on electrostatic potential (ESP) fitting.
The calculation of the Coulomb coupling with transi-
tion ESP charges (TrEsp) was developed by Renger and
coworkers (Madjet et al., 2006), and represents now a
widely used method (Kenny and Kassal, 2016; Olbrich
and Kleinekathofer, 2010; van der Vegte et al., 2015).
The Coulomb coupling can also be obtained by directly
evaluating the integral, Eq. (28). The first calculation of
this kind was developed by using a numerical integration
over a three-dimensional grid (Krueger et al., 1998). This
numerical approach is called the Transition Density Cube
(TDC) method, and in principle gives the exact Coulomb
contribution of the electronic coupling with an appropri-
ately chosen grid. The TDC method has been employed
extensively (Bricker and Lo, 2014, 2015; Scholes, 2003;
Scholes and Fleming, 2000; Scholes et al., 1999). How-
ever, more reliable and efficient approach is to perform
the integration of Eq. (28) analytically. This also allows
the inclusion of explicit screening of the Coulomb inter-
action due to the environment (Curutchet et al., 2009;
Hsu et al., 2001a; Iozzi et al., 2004).
The coupling is also significantly affected by the
presence of the environment surrounding the pigments.
While for site energies, environment effects are generally
seen as a shift depending on the nature of the transi-
tion and the characteristic of the environment, the elec-
tronic couplings are affected through two mechanisms,
each leading to a specific change.
First of all, the environment can change both the ge-
ometrical and the electronic structure of the pigment
molecules and modify their transition properties,
i.e.
transition dipoles and transition densities. These changes
will be "implicitly" reflected in a change of the Coulomb
coupling which will be generally enhanced. The second
effect is due to the polarizable nature of the environment
which acts as a mediator of the pigments' excitations.
The resulting "explicit" effect acts to reduce the magni-
tude of the direct (Coulomb) coupling. For this reason it
is common to say that the coupling is "screened" by the
environment.
jk, where J c
The simplest way to account for the explicit effect
of the environment is to introduce a screening factor s
such that Jjk = sJ c
jk is the direct electronic
coupling between chromophores with the implicit effect
taken into consideration. This definition is in line with
the standard definition of dielectric screening by environ-
ment in electromagnetism but may not always be easy to
determine. In practice, one may introduce a total screen-
ing factor st such that Jjk = stJ c,0
jk is the
bare electronic coupling between chromophores without
including the implicit effect. The difference between s
and st can be significant, and care should be taken in
jk , where J c,0
using the proper definition. More discussion of this issue
will be provided later.
The screening factor s can be clearly related to the
inverse of an effective dielectric constant ∞ which repre-
sents the electronic response of the environment approxi-
mated as a dielectric. ∞ is generally approximated with
the square of the refractive index of the medium: when
the environment is characterized by a refractive index of
1.4 (typical for hydrophobic region of protein environ-
ments) the screening factor is ∼ 0.5 and the electronic
coupling is reduced by a factor ∼ 4. Within this approx-
imation, the screening does not depend on the interact-
ing chromophores, neither on their relative orientation
and distance. Moreover, at short distances, the dielectric
medium can be excluded from the intermolecular region,
leading to more complex effects. In particular cases, this
can also enhance the coupling.
To achieve a more accurate modeling of the environ-
mental effects on electronic couplings, a combination of
a quantum mechanical description of the pigments with
a continuum model can be introduced. Various formula-
tions of continuum solvation models (Tomasi et al., 2005)
can be used to this end. By applying the polarizable con-
tinuum model (PCM) (Mennucci, 2012) described in Ap-
pendix B, the response of the environment (i.e. the polar-
ization of the dielectric) is described by a set of induced
(or apparent) charges spreading on the surface of the
molecular cavity embedding the chromophores. Within
this framework, J r
jk becomes:
J r,PCM
jk
=
drρtr∗
j
(r)
1
r − rt
qt(∞; ρtr
k )
(31)
(cid:20)(cid:90)
(cid:88)
t
(cid:21)
jk)/J c
jk +J r
Conceptually,
the electronic transition in the chro-
mophore k drives a response in the polarizable medium,
which in turn, affects the transition in the chromophore
j. It is important to note that the induced charges qt are
calculated using the optical permittivity of the medium,
in order to account for the fact that only the electronic
component of the polarization can respond. Also in this
case, we can define an effective screening factor as the ra-
tio between the total coupling and the direct interaction,
namely s = (J c
jk. A QM/PCM study has been
conducted to investigate the dependence of the screening
factor on the nature of the interacting chromophores and
their relative arrangement, using different pairs (chloro-
phylls, carotenoids and bilins) extracted from LHCs (Sc-
holes et al., 2007). It has been shown that at large inter-
chromophore separation ( > 2 nm ) the screening factor
is practically constant. At closer distances, instead, the
screening shows an exponential behavior approaching to
∼ 1 for distances of the order of few tenths of nanome-
ter. At such close distance, the chromophores share a
common cavity in the medium (the environment cannot
access the region where the two chromophores are at close
contact) and the screening effect is reduced.
18
jk/J c,0
As noted above, the scaling factor s accounts for
only the explicit screening. On the other hand, the
environment can also affect the transition densities of
chromophores, causing implicit change of electronic cou-
plings, which we call here si. A detailed analysis of
si = J c
jk was reported in detail in a follow-up paper
(Curutchet et al., 2007), which showed that it is not pos-
sible to define a similar empirical expression for si as has
been done for s. The total empirical screening factor we
defined earlier is given by st = s· si. In general, it cannot
be fitted by a simple function of the distance. Thus, care
should be taken in estimating the distance dependence
of the total screening factor.
An alternative way to account for environment effect
in the coupling, is through the so-called Poisson-TrEsp
method developed by Renger an coworkers (Adolphs
et al., 2008; Renger and Muh, 2011). The TrEsp charges
of the pigments are placed in molecule-shaped cavities
that are surrounded by a homogeneous dielectric with a
dielectric constant, which represents the optical permit-
tivity of the protein and solvent environments. A Poisson
equation is then solved for the electrostatic potential of
the TrEsp of each pigment and the resulting potential is
finally used to calculate the coupling.
Despite the success of these continuum approaches, an
atomistic description of the protein environment is ex-
pected to give a more complete description.
In those
cases, in fact, the dielectric response varies locally and
specific interactions between the chromophores and the
protein can be established. A classical formulation can
still be used by introducing a Molecular Mechanics (MM)
description in which each atom of the environment is rep-
resented by a fixed point charge to mimic the electrostatic
effects. In order to properly account for all the environ-
mental effects, however, the MM model has also to be po-
larizable. Various strategies are possible (see Appendix
B). In the context of LHCs, the most used approach is
represented by the induced dipole (ID) formulation where
an atomic polarizability is added to the fixed charge to
describe each atom of the environment. Within this MM-
Pol framework, the J r
jk term becomes (Curutchet et al.,
2009):
(cid:20)(cid:90)
= −(cid:88)
p
J r,MMPol
jk
(cid:21)
drρtr∗
j
(r)
(rp − r)
rp − r3
µp(ρtr
k )
(32)
where the transition densities ρtr
k induce a response in the
environment which is represented by the induced dipoles
µp. As in the case of the QM/PCM, ρtr
k here is also
calculated self consistently with the polarization of the
environment. A mixed continuum/atomistic strategy has
also been proposed (Caprasecca et al., 2012): in this case,
jk is the sum of J r,PCM
J r
, and both terms
are obtained in a fully polarizable scheme.
and J r,MMPol
jk
jk
As a cautionary remark, we note that, in the case of
any atomistic description, the QM-environment interac-
tions, particularly those at short range, depend critically
on the configuration of the environment. Therefore, sev-
eral configurations of the whole system need to be taken
into consideration to get a correct sampling. Commonly,
the sampling is obtained by using classical Molecular Dy-
namics (MD). This sampling is not needed when a con-
tinuum approach is employed since it implicitly gives a
configurationally averaged effect due to the use of macro-
scopic properties.
A completely alternative way for the calculation of the
electronic coupling between two sites is that based on a
full quantum calculation on the dimeric unit, which is
also known as super-molecule approach. These schemes
are based on the diabatization of the electronic Hamilto-
nian of the whole donor-acceptor system, and in principle
yield the "exact" coupling including exchange and over-
lap interactions within the electronic structure method
used (Hsu, 2009; You and Hsu, 2014).
Consider a molecular system composed of two moi-
eties, which can be either two separate molecules or two
fragments of the same molecule. An electronic structure
calculation on the entire system necessarily yields the adi-
abatic states, which are the eigenstates of the electronic
Hamiltonian at a specific nuclear geometry. However, a
diabatic picture better describes the states involved in en-
ergy transfer. Within this picture, the electronic Hamil-
tonian is written in a basis of localized states, and is not
diagonal. For example, considering only two states, the
Hamiltonian matrix reads as follows:
(cid:18) Ej Jjk
(cid:19)
Jjk Ek
Hel =
,
(33)
where Ej and Ek are the energies of the diabatic states,
and Jjk is the electronic coupling between those states.
At the avoided crossing point, the condition Ej = Ek
means that the energy gap between the adiabatic states
is 2Jjk. Therefore, Jjk may be computed as half of the
energy gap. Obviously, this condition holds at all ge-
ometries when the two fragments are identical. In more
general cases, a localization scheme is needed to define
the diabatic states, and to find the transformation matrix
between adiabatic and diabatic bases.
There is no unique choice for the diabatic states, the
choice of which also can alter the definitions of Coulomb
and exchange interactions (Vura-Weis et al., 2010). Sev-
eral schemes have been proposed, such as the fragment
excitation difference (FED) (Hsu et al., 2008) or the
more recent fragment transition difference (FTD) scheme
(Voityuk, 2014). Within this framework, an additional
operator (which we shall call Y ), representing an observ-
able that has its extrema in the diabatic states, is intro-
duced; Y can be defined in such a way that it has eigen-
values 0 and ±1. The localized states are therefore those
states that diagonalize Y . The eigenvectors of the matrix
Y form the unitary matrix transformation U (i.e., U†YU
is diagonal), which is the adiabatic-to-diabatic (ATD)
19
transformation, within the assumption that the two adi-
abatic states are a linear combination of the localized
states of interest. By applying the same transformation
to the diagonal energy matrix, one can obtain the Hamil-
tonian in the diabatic basis, and the electronic coupling
Jjk as the off-diagonal elements. Namely,
Jjk = (Ek − Ej)
(cid:113)
Yjk
(Yj − Yk)2 + 4Y 2
jk
.
(34)
Applying the transformation U to the adiabatic states
yields states that are localized as much as possible, and
similar to the initial and final states. However, in many
cases, a 2-state adiabatic basis is not sufficient to re-
trieve completely localized states. In fact, an adiabatic
state could be a combination of many diabatic states
of both the donor and the acceptor. Moreover, charge-
transfer states can mix with excitonic states, and vice-
versa (Voityuk, 2013; Yang and Hsu, 2013; You and Hsu,
2014). Another issue with the 2-state model is its in-
ability to compute the couplings between states that are
more weakly coupled. To overcome these limits, one can
resort to a multi-state formulation. The generalization
to a multi-state model is not straightforward, as the ad-
ditional operators only have three different eigenvalues
(0 or ±1). This means that the diagonalization of these
operators will only separate the adiabatic basis in three
subspaces, and therefore there is no unique choice for
the transformation U from the adiabatic to the diabatic
basis.
C. Applications of electronic structure calculations to LHCs
We present here a summary of three computational
studies aimed at understanding the molecular origin of
the protein tuning of the excitonic properties of three
LHCs introduced earlier, the FMO complex of green sul-
fur bacteria, LH2 complex of purple bacteria, and phy-
coerythrin PE545 of cryptophyte algae.
1. FMO complex
As mentioned in the overview of Sec. II, the spectral
features in FMO arise from the subtle tuning of the indi-
vidual site energies due to the surrounding protein envi-
ronment. To shed light on this effect, the environment-
induced changes in the site energies were analyzed in de-
tail using the combination of TDDFT and different clas-
sical models based either on the atomistic or continuum
descriptions (Jurinovich et al., 2014).
The trimeric crystal structure of Prosthecochloris aes-
tuarii (pdb entry 3EOJ, res. 1.30 A), which contains
eight BChl per monomer, was used (Tronrud et al., 2009).
Considering the C3 symmetry of the system, the QM
analysis was performed only on the monomer system
shown in Figure 3. The corresponding excitation ener-
gies of BChl were computed at the TD-B3LYP/6-31G(d)
level of theory, the results of which are shown in Fig. 9.
FIG. 9 Site energies (eV) of the eight BChls computed on the
crystal structure with three different models: the Full model
(full line and circles) includes all BChl atoms, the Truncated
model (dotted line and triangles) does not include the atoms
of the phythyl chain, the Tail@MMPol model (diamonds) de-
scribes the atoms of the phythyl chain as MMPol sites. Inset:
Structure of BChl 1 for which the chain is folded in a confor-
mation capable of interacting with the porphyrin ring.
In FMO complex, like in other LHCs, each pigment
is confined in a specific binding pocket, surrounded by
the protein matrix, which may perturb the pigment's ex-
citation through two distinct effects: 1) a direct effect
on the electronic states determined by pigment-protein
interactions; 2) an indirect effect due to the modifica-
tions induced by the environment on the geometry of the
pigment. The effect of this can be seen clearly in the
results of calculations: if the direct environmental effects
are switched off, the resulting site energies are still differ-
ent for the different BChls due to the indirect effects on
geometries induced by the local environment. In particu-
lar, the isolated BChl has a planar equilibrium structure
(excluding the flexible phythyl chain). However, when
embedded in the protein environment, such planarity be-
comes perturbed differently in different binding pockets.
The phythyl chain of BChl also plays a role in differ-
entiating the excitation energy through variation in its
conformations. According to the crystal structure, all
the chains adopt an outstretched configuration, except
for BChl 1 and 4, for which the chain is folded in a con-
formation capable of interacting with the porphyrin ring
(see Figure 9). To investigate these effects, site energy
calculations were performed using the crystallographic
20
structure of the pigments and 1) including all the BChl
atoms at full quantum mechanical level calculations, 2)
replacing the phythyl chain with a methyl group but
still treated at the full quantum level, and 3) replacing
the entire phythyl chain with classical polarizable MM
sites through the MMPol approach (tail@MMPol).
In
this last model, the boundary between the QM and the
classical subsystem was treated by using the link atom
scheme. The results shown in Figure 9 reveal that, when
the phythyl chain is close to the porphyrin ring (BChl
1 and 4), the site energy is lowered, as if it introduces
an "additional environmental" effect not present in the
other pigments. This effect seems to be largely due to
a classical electrostatic and polarization interaction. In
fact, when the chain is treated at MMPol level, a similar
lowering of the energy is observed, albeit not as large as
in the full QM description.
An important question concerning the role of protein
environments in tuning the excitation energies of pigment
molecules is the relative importance of specific molecu-
lar level interactions versus mean-field (or bulk) effects.
This issue can be analyzed through detailed and careful
analysis of the two contributions. As an example, val-
ues of excitation energies for 1 through 7 BChl molecules
obtained either with the atomistic (MMPol) or the con-
tinuum (PCM) model are compared in Figure 10. BChl
8 is excluded from this analysis because of the exter-
nal location of its binding site, being more exposed to
the solvent. This is manifested in drastic difference be-
tween PCM and MMPol descriptions of the environment,
in particular when the MMPol model is based on the
crystal structure and does not contain surrounding wa-
ter molecules.
FIG. 10 Site energies (eV) for the 1-7 BChls computed on
the crystal structure including the environment effects at
PCM (dotted line) and MMPol level (full line). A third
(me)QM/PCM model combining PCM with a QM descrip-
tion of the interacting residues is also shown (dashed line).
As described in the previous section, the PCM model
describes the bulk effect of the environment through an
effective dielectric constant, which corresponds to a re-
sponse averaged over many different configurations of the
environment.
It is evident that when some specific or
directional interactions are present (i.e. H-bond or co-
ordination bonds), their effects cannot be accounted for
in such an "average over the medium" model. A practi-
cal solution for this deficiency is to enlarge the definition
of the quantum mechanical subsystem being in the di-
electric medium by including not only the pigment but
also its proximate protein residues. Through this mini-
mal environment (so called "me") approach, the effects of
short-range interactions can be taken into account within
the PCM model.
The
between
comparison
QM/PCM and
(me)QM/PCM shows very similar energies for BChls 2
and 5. These two BChls have weaker axial interactions
(a water molecule for BChl 2 and a carbonyl group
belonging to the protein backbone for BChl 5) with
respect to the other pigments, which are coordinated
with the nitrogen of a histidine residue.
In the lat-
ter cases, the site energies are red-shifted by about
160 cm−1 when including the QM residues. In principle,
the QM/MMPol description should be able to describe
both the bulk effect of the protein and the specific
electrostatic interactions of the closer residues. Indeed,
the QM/MMPol and the (me)QM/PCM site energies
are in good agreement for all pigments except for BChl
6.
It has to be also noted that, for such a pigment, a
positively charged residue (an arginine) directly interacts
with the side group of the porphyrin ring. In this case,
the description of this residue at a QM level allows
to include non classical interactions (such as possible
charge transfers) between the BChl and the protein,
which can significantly affect the electronic excitation.
In summary, the inclusion of the interaction with the
protein environment results in a redshift of the excita-
tion energies of BChls in the FMO complex, but in ways
depending on different local environments. Thus, a sim-
ple continuum model alone is not able to properly ac-
count for both short and long-range environmental ef-
fects. This point is also supported by a detailed electro-
static analysis of the site energy funnel in FMO (Muh
et al., 2007), which showed that the electric field from
the α-helices defines the direction of excitation energy
flow in the FMO protein, whereas the effects of amino
acid side chains largely compensate each other. Thus,
it is important to have a complete and balanced picture
of the fine tuning by protein environments in order to
characterize the FMO complex properly. An effective
and still practical way for this is to include the residues
more strongly interacting with the pigments in the defi-
nition of the QM subsystem, while keeping a continuum
description for the rest. Nonclassical short-range effects
can in fact have a non-negligible contribution to the pat-
21
tern of BChl excitation energies in the FMO complex.
In addition, the BChl phythyl chain can play a subtle
but significant role, given the small differences found in
the energies of the pigments. In particular, the largest
effects are found when a folded conformation close to the
porphyrin ring is possible, such as for BChl 1 and 4. The
effects of the tail can be explained as a combination of
overlap interaction at the quantum level and electrostatic
plus polarization at purely classical level. It is worthwhile
to note that the latter components can be recovered by
using an MMPol description of the chain.
2. LH2
The early atomistic level calculations (Hsu et al.,
2001b; Hu et al., 2002, 1997; Krueger et al., 1998) were
based on the X-ray crystal structures (Koepke et al.,
1996; McDermott et al., 1995) and focused on calculating
electronic coupling constants between excitations. The
outcomes of these calculations, in combination with em-
pirical correction factors to complement numerical errors,
laid the foundation for many exciton models that fol-
lowed.
In addition, already in 2002, all atomistic MD
simulation of LH2 complex was shown to be feasible
(Damjanovi´c et al., 2002), providing the basis for more
recent advances (Cupellini et al., 2016; Jang et al., 2015;
Montemayor et al., 2018; Olbrich and Kleinekathofer,
2010; Sisto et al., 2017). In this section, some of the out-
comes of these recent computational studies for the LH2
antenna complex of Rps. acidophila are summarized.
QM/MM calculations (Cupellini et al., 2016) have
been performed using (i) the crystal structure resolved at
2.0 A (PDB code: 1NKZ) (Papiz et al., 2003) and (ii) the
configurations extracted from a room-temperature sam-
pling, carried out through an MD simulation of the LH2
complex within a solvated lipid membrane. Environmen-
tal effects on the calculations of the excitonic parameters
have been introduced in terms of the polarizable (MM-
Pol) embedding combined with a TD-DFT description
based on the CAM-B3LYP functional and the 6-31G(d)
basis set. The comparison of a "static" description using
a single crystal structure with a "dynamic" one includ-
ing structural and electronic fluctuations of the pigments
coupled to electrostatic and polarization fluctuations of
the environment, allows detailed and molecular-level in-
vestigation of the key factors contributing to the charac-
teristics of excitons and their changes in temperature.
For the case of the dynamic model, a different Hamil-
tonian matrix was calculated for each of the 88 config-
urations extracted from the MD simulation and the re-
sulting "instantaneous" excitonic parameters were finally
averaged to be compared with those obtained on the sin-
gle crystal structure. Indeed, the two sets of data show
significant differences. The B850 exciton splitting de-
creases from 1517 to 1275 cm−1 (-16%) moving from the
static description based on the crystal structure to that
based on the MD simulation, which is consistent with
experimental variation between low and room tempera-
ture Davydov splittings (-13%) (Pajusalu et al., 2011b).
This behavior is also reflected on the absorption spec-
trum, where the B850 band blue shifts when the tem-
perature is increased, whereas the B800 peak does not
shift (Trinkunas et al., 2012). These results suggest that
the B850 blue shift is mainly due to a reduction of the
inter-pigment couplings. In addition, the site energy dif-
ference between B800 and B850 pigments is also reduced
on average by ∼50 cm−1 in the calculation based on the
MD simulation.
The static and the dynamic models also give different
descriptions of the environment effects. Using the crys-
tal structure, the inclusion of the MMPol environment
causes a redshift of the BChls' excitations of ∼880 cm−1,
with small differences among the different pigment types
(< 25 cm−1). A red shift is also obtained for the struc-
tures coming from the MD trajectory. However, in this
case, the shift in the excitation energies of α and β BChls
of the B850 unit is about 830 cm−1, whereas the shift for
B800 is only 610 cm−1. When considered within the ap-
proximation of transition dipole moments, an interesting
environment effect can be observed with respect to their
orientations. In both models, the environment does not
affect the out-of-plane tilt, but it has a non-negligible in-
fluence on the in-plane tilt. The changes are almost the
same for α and β BChls of the B850 unit.
The different features that can be found between the
two static and dynamic models are well reflected in the
exciton delocalization. There are various measures of ex-
citon delocalization (Dahlbom et al., 2001; Jang et al.,
2001), and a well-known measure is the following inverse
participation ratio (IPR):
(cid:34)
(cid:104)(cid:88)
(cid:35)−1
L
IPR
D =
(Ujk)4(cid:105)
(35)
k
IPR
where Ujk is the unitary transformation as defined above
Eq. (3), j is the index for each exciton state, k is the in-
dex for each site excitation state, and (cid:104)···(cid:105) represents
averaging over all exciton states with proper thermal
D ranges from 1, for a fully
weights. By definition, L
localized state, to the number of total sites for a com-
pletely delocalized state.
Due to the CN symmetry (with N = 8−10) of the LH2
complex, in the absence of disorder, the exciton states are
all doubly degenerate with the exception of the lowest
energy exciton states for both the B850 and B800 units
(plus the highest exciton state for B850 with odd N ). In
such a perfectly symmetric arrangement for N = 9, for
example, the delocalization lengths are maximum 18 (9)
for the non degenerate states of the B850 (B800) unit,
and 12 (6) for the doubly degenerate states. The cal-
culations in the static model indeed give delocalization
22
lengths which are close to the maximum values, even in
the B800 unit, despite small electronic couplings between
BChls. Instead, adding disorder in the BChls' site ener-
gies through the dynamic model results in a localization
of the excitons with a general reduction of L
D . In the
B850 unit, the large couplings allow the exciton to remain
delocalized even in the presence of disorder but with a
significant reduction of L
D to an average value of typi-
cally 8±2. On the contrary, in the B800 ring the average
delocalization length reduces to ∼ 1.4.
IPR
IPR
Although LIPR
D is a well established measure of delocal-
ization, its physical implication in the partially delocal-
ized regime is not always clear. For this reason, different
measures of delocalization length have been tested, which
have different degrees of sensitivity (Dahlbom et al.,
2001; Jang et al., 2001). For example, a simple alter-
native measure can be defined as follows:
jk}(cid:105) .
D = (cid:104)(cid:88)
min{1, NcU 2
(36)
JDS
L
k
IPR
This measure does not give a weight more than one for
D while approaching the same values
each site unlike L
in both full localized and delocalized limits. For the B850
unit (Jang et al., 2001), this measure was shown to be
D to the nature of the disorder and
more sensitive than L
also results in larger estimate of delocalization length.
IPR
The computational results obtained for both the single
crystal structure and MD configuration can also be used
to simulate the circular dichroism (CD) spectrum of LH2.
The CD spectrum represents a fingerprint that is unique
for each LHC as it is determined by the nature of the
excitons and the geometrical characteristics of the aggre-
gate of pigments. Figure 11 compares the two calculated
CD spectra with those measured at low (77K) and room
temperature.
In the case of the dynamic model based
on the MD simulation, the spectrum reported in Fig. 11
is the average of the instantaneous spectra obtained for
each of the 88 configurations.
The experimental CD spectrum is characterized by two
"couplets" corresponding to the excitonic signal due to
the B800 and B850 ring, respectively. Due to the in-
crease of the temperature (from 77 K to room temper-
ature) both the broadening and the relative intensities
of the couplets change. In particular, the B800 positive
band almost vanishes. The CD spectrum is sensitive to
the inter-ring coupling, and the B850 couplet borrows in-
tensity from the B800 band. The small B800–B850 mix-
ing happens despite the site energy differences and the
static disorder, and breaks the symmetry of the B800 and
B850 couplets. The only way to reproduce the asymmet-
rical B800 couplet is to consider some B800–B850 mixing.
Unlike the spectrum obtained from the crystal structure,
here the B800 couplet amplitude is nearly half of the
B850 couplet. This can be explained as an effect due to
the disorder in excitation energy of each BChl, combined
with the small intra–B800 coupling.
23
ble multi-configurational character of ground and excited
states of the pigments can lead to new insights with re-
spect to DFT-based investigations, as it has been shown
by recent ab initio multi-reference calculations on bac-
teriochlorophylls and carotenoids of LH2 (Anda et al.,
2016; Segatta et al., 2017).
3. PE545
The effect of the protein (and the surrounding sol-
vent) on the excitonic properties of PE545 has been in-
vestigated through QM/MMPol calculations for a series
of structures extracted from a classical MD simulation
of the LHC in water at room temperature (Curutchet
et al., 2011b). The QM/MMPol results were comple-
mented with those based on a continuum (PCM) descrip-
tion of the protein and water environment. The compar-
ison between the two sets of calculations can in fact be
used to show how the heterogeneous properties of the
protein can modify the local screening of the electronic
couplings between the bound pigments. MMPol results
were averaged over 140 configurations of PE545 and wa-
ter molecules extracted from the MD simulation, whereas
PCM results were obtained using the cluster of the pig-
ments as found in the ultrahigh resolution crystal struc-
ture of the complex. In all cases, the first low-lying π−π∗
excited state of the 8 bilins were computed together with
all the corresponding electronic couplings between them.
The full transition densities of the pigments were used for
these calculations employing Eqs. (28), (31), and (32). A
comprehensive set of calculation results performed at the
CIS/6-31G level are available (Curutchet et al., 2011b).
The heterogeneous nature of the protein environment
can significantly modulate the coupling through both
changes in the transition density of each pigment and in
the extent of screening of the corresponding interactions.
In the QM/MMPol approach, both effects are included in
the coupling term by considering both interchromophore
distances and orientations, as well as the heterogeneous
polarization of the protein and solvent environment. For
more quantitative understanding, it is useful to compare
the screening factor (and a related effective permittivity)
defined differently for each pair j and k as follows:
sjk =
1
jk
=
jk + J r,MMPol
J c
jk
J c
jk
,
(37)
where J c
jk is the Coulomb interaction given by Eq. (28),
which includes the implicit effect of the environment on
transition densities, whereas J r,MMPol
is the explicit MM-
Pol term as described in Eq. (32).
jk
As illustrated in Figure 12, PCM values of jk are sim-
ilar for all pairs while MMPol values present a significant
spread (from 2.6 to 1.2). On the other hand, PCM and
MMPol values become very similar when averaged over
FIG. 11 Upper Panel: The crystal structure used in the
static model (left) and examples of configurations extracted
from MD used in the dynamic model (right) of LH2. Lower
panel:
(left) measured CD spectra at 77 K (dotted line)
and room temperature (full
line) and (right) calculated
TDDFT/MMPol spectra using the static (dotted line) and
the dynamic model (full line).
The modeling of the CD spectra suggests the static
and the dynamic models can be effectively used to rep-
resent LH2 in two different situations, namely, at zero-
temperature limit and at room temperature. The main
features of the CD spectra and their temperature depen-
dence are in fact well reproduced. Moreover, this analysis
indicates that the difference in the exciton structure of
LH2 at low and room temperatures are mainly related to
fluctuations in the relative orientations of the BChls (and
of the corresponding couplings), rather than changes in
the ring size as was previously suggested (Pajusalu et al.,
2011b).
Despite impressive advances as described above and
recent demonstration of the capability of on the fly cal-
culations (Sisto et al., 2017), further improvements in
accuracy and efficiency are still needed. Some of key
issues to be addressed in this respect include the differ-
ences in the excitation energies of α-, β-, and γ-BChls,
solvatochromic shift due to hydrogen bonding, and con-
tribution of carotenoids. Although some of the early ex-
citon models based on the X-ray crystal structure sug-
gested that (Rancova et al., 2012) α- and β-BChls have
about 300 cm−1 difference in their excitation energies,
QM/MM calculations for configurations extracted from
MD simulation showed that the two energies are virtually
the same in actual protein environments (Cupellini et al.,
2016). Combination of MD simulation and DFT cal-
culations also demonstrated that hydrogen bonding can
cause red shift of excitation energies by about 500 cm−1
or larger (Jang et al., 2015; Montemayor et al., 2018).
Moreover, accounting for the contribution of the possi-
all pairs. The agreement between averaged results can be
explained through a careful analysis of the components of
each model. In the continuum method, the screening is
described in terms of a set of induced charges spreading
on the surfaces of the cavities embedding the pigments.
These charges represent the polarization of the environ-
ment induced by the electronic transition in the donor.
These are calculated in terms of the optical component
of the dielectric permittivity used to represent the mixed
protein-water environment (namely 2.0). In the MMPol
approach, instead, the screening is calculated in terms
of induced dipoles originating from the electronic transi-
tion in the donor. Within this description, the induced
dipoles are determined by the atomic polarizabilities used
to mimic the protein (and the water) atoms. The latter,
when used to simulate the macroscopic polarization of
the whole environment, gives an effective permittivity of
2.3, which is very close to that used in the PCM model.
The similarity between the continuum and the atomistic
model, however, disappears when the pigment pairs are
analyzed separately. Each of them in fact is expected to
feel a different local environment determined by the vari-
ety of residues, respectively presenting a different degree
of polarization.
FIG. 12 Effective dielectric constant corresponding to each
pair (ik ) reported as a function of the interchromophore dis-
tance: circles refer to the average QM/MMPol@MD data and
squares to the QM/PCM calculations on the crystal struc-
ture. The dotted line indicates the average value obtained
on all MMPol@MD values. The two insets indicate the two
pairs showing the smallest (DBV19A − DBV19B) and largest
(PEB50/61C − PEB50/61D) ik value.
The smallest effective permittivity (1.35) is experi-
enced by the peripheral DBV19A-DBV19B pair. On the
contrary, for the central PEB50/61C − PEB50/61D pair,
the coupling is significantly more attenuated by the pro-
tein than in other pairs and the resulting effective permit-
tivity is the highest (2.57). To understand the origin of
these differences, the screening obtained by the MMPol
approach can be dissected into contributions arising from
the different protein residues and waters. Examination
of the central pair shows that all the polypeptide chains,
together with waters, act to reduce the interaction, re-
24
sulting in a large screening effect. A completely different
picture appears for the peripheral pair. In this case, it
seems that the protein is organized in such a way as to
enhance the electronic coupling while further effect of the
surrounding water adds a strong screening contribution.
D. Bath spectral density
As represented by Eq. (23), the contribution to the
bath modes for a molecule embedded in a flexible en-
vironment can be classified into two different sources,
one from the internal vibrations of the molecule and the
other from the motions of the molecule within the envi-
ronment. Conventionally, it is assumed that the internal
vibrations constitute only high-frequency underdamped
modes, which are reflected in sharp peaks of the spec-
tral density. The low frequency intermolecular modes are
often viewed as being induced entirely by the surround-
ing environment, resulting in a continuous contribution
to the spectral density. However, for the case of pig-
ment molecules in LHCs, intramolecular vibrations are
also shown to make significant contributions to the low
frequency modes.
Accurate determination of spectral densities Jjk(ω) of
Eq. (22) is a difficult task in general, let alone confirm-
ing the validity of the form of the HLHC as assumed in
Eq.
(21). For the case where there is only one pig-
ment molecule, the corresponding spectral density can
be determined experimentally using either spectral hole
burning (SHB) or fluorescence line narrowing (FLN). The
former removes inhomogeneity by burning specific zero-
phonon transition and the latter by obtaining fluores-
cence lineshape following excitation at the red edge of
the ensemble absorption lineshape. Model spectral den-
sities fitted to these experimental ones have been used
for the construction of EBHs.
A popular example of the model bath spectral density
is the following Ohmic spectral density with Drude cutoff:
JDrude (ω) = 2λ
ω/ωc
1 + (ω/ωc)2 .
(38)
For this spectral density, the well-known choice of pa-
rameters for the FMO complex is λ = 35 cm−1 and
ωc = 106 cm−1. Although not as widely as in the case
of the FMO complex, this spectral density was also used
for the LH2 complex (Chen et al., 2009; Meier et al.,
1997a; Zhang et al., 1998) with λ = 240 cm−1 and
ωc = 40.8 cm−1 based on the fitting (Meier et al., 1997a)
of photon-echo data (Jimenez et al., 1997). Another
well-known form is the Ohmic or super-Ohmic spectral
density with exponential cutoff. A combination of these
forms were constructed for the B850 unit of the LH2 com-
plex based on the fitting of fluorescence-line narrowing
data as follows:
JJN S (ω) = π
(cid:18)
γ1ωe−ω/ωc,1 + γ2
e−ω/ωc,2
ω2
ωc,2
(cid:33)
+γ3
ω3
ω2
c,3
e−ω/ωc,3
,
(39)
with γ1 = 0.22, γ2 = 0.78, γ3 = 0.31, ωc,1 = 170 cm−1,
ωc,2 = 34 cm−1, and ωc,3 = 69 cm−1. The above spec-
tral density (or simpler version with only Ohmic term)
was used for the modeling of low temperature SMS line-
shapes (Jang and Silbey, 2003a; Jang et al., 2011; Kumar
and Jang, 2013) and calculation of exciton transfer rates
(Jang et al., 2014, 2004, 2007).
Recently, more refined experimental technique called
∆-FLN, which combines the merits of SHB and FNL,
was developed (Ratsep and Freiberg, 2007; Timpmann
et al., 2000a) and applied to various LHCs (Kell et al.,
2013; Pieper et al., 2011; Ratsep et al., 2008). In particu-
lar, a recent analysis (Kell et al., 2013) suggested that low
frequency region of the model Ohmic spectral density is
not consistent with the experimental lineshape, propos-
ing a new form with log-normal distribution in the low
frequency limit. While this exposes a potential deficiency
of conventional model spectral densities that typically as-
sume algebraic behavior in the small frequency limit, it
is not yet clear whether such log-normal distribution is
an apparent effect of residual degree of inhomogeneity
and anharmonic contribution of the bath. In addition,
for general LHCs with more than one pigment molecules
with similar excitation energies, it is not yet clear how
even the ∆-FLN approach can accurately determine site-
specific spectral densities represented by Eq. (22).
Computational determination of the spectral densities
are feasible, but are also beset with a few theoretical
and practical issues. A well-known approach is to use
a mixed quantum-classical approach assuming that the
bath is classical, and to calculate the following bath en-
ergy gap correlation function through classical MD simu-
lation (Aghtar et al., 2013; Chandrasekaran et al., 2015;
Olbrich et al., 2011b; Shim et al., 2012; Viani et al., 2014;
Zwier et al., 2007):
Ccl
E,jk(t) = (cid:104)∆Ej(t)∆Ek(0)(cid:105)cl ,
(40)
where ∆Ej(t) is the difference of the bath energy for sj(cid:105),
which is created at time t = 0, and that for the ground
electronic state. Assuming that this is related to the real
part of the quantum correlation function of Bjk defined
in Eq. (20), one can calculate the spectral density using
the following relation (Olbrich et al., 2011b):
Jjk(ω) =
2
tanh(
βω
2
)
dt Ccl
E,jk(t) cos(ωt) .
(41)
(cid:90) ∞
0
On the other hand, if we rely on the approximation of Eq.
(21) and introduce mass-weighted normal mode Qn such
√
2ω3/2
n gj,nQn, it is straightfor-
25
gj,ngk,nω3
n(cid:104)Qn(t)Qn(0)(cid:105)cl
dω Jjk(ω)
cos(ωt)
ω
.
(42)
that ωjgj,n(bn + b†
ward to show that
Ccl
n) =
E,jk(t) = 2(cid:88)
(cid:90) ∞
(cid:90) ∞
2
πβ
Jjk(ω) = βω
n
0
=
0
Then, through cosine transformation,
dt Ccl
E,jk(t) cos(ωt) .
(43)
While this becomes equivalent to Eq. (41) if the vibra-
tional quanta of all the bath modes are smaller than ther-
mal energy, the discrepancy between the two becomes
substantial for high frequency vibrational modes. Given
that harmonic oscillator approximation is more appropri-
ate for high frequency modes, which result mostly from
intramolecular vibration, Eq. (43), known as harmonic
approximation, is considered more reliable than Eq. (41)
in general (Chandrasekaran et al., 2017; Valleau et al.,
2012). For the bath modes coming from protein envi-
ronments, this may not be necessarily true because of
anharmonic and nonlinear effects.
Another important issue is that a long time window
is required in order to achieve a complete sampling for
low-frequency vibrations. For example, a vibration at
∼10 cm−1 has a period of ∼3 ps, requiring a sampling
time window of at least 100 ps. Conversely, to sample in-
tramolecular modes, one needs a rather short time step,
i.e. ≤ 5 fs (Chandrasekaran et al., 2015; Shim et al.,
2012; Valleau et al., 2012). These two opposing factors
make the mixed approach computationally very expen-
sive, requiring tenths of thousands of electronic structure
calculations for a single spectral density.
The most concerning issue of the mixed quantum-
classical approach is the quality of the MD trajectory
involved. In fact, an assumption of the spectral density
approach is that the relevant dynamics come from the
evolution of the energy gap dictated by the ground-state
Hamiltonian. However, the dimensions of the systems
under study and the time scales involved compel the nu-
clear trajectory to be computed with a classical force
field, i.e., using fitted parameters, whereas the energy
gap is calculated with a quantum mechanical Hamilto-
nian. The Hamiltonian used to propagate the nuclear
positions may be critically different from the one used
in the calculation of the energy gap, leading to an effec-
tive excited-state PES that is completely different from
the "real" quantum mechanical one. The ground-state
PES could differ in equilibrium position, normal modes,
and frequencies. All of these factors contribute to the
shape of the spectral density (Kim et al., 2015; Lee et al.,
2016; Zwier et al., 2007). Indeed, it has been observed
that the resulting spectral density is strongly dependent
on the force field parameters, rather than on the quan-
tum mechanical Hamiltonian (Aghtar et al., 2013; Chan-
drasekaran et al., 2015; Wang et al., 2015). For this
reason, in order to effectively use the mixed quantum-
classical method, one should carefully assess the quality
of the force field parameters by comparing them to the
quantum-mechanically calculated PES, and possibly de-
velop ad-hoc parameters specifically designed for subse-
quent excitation energy calculations (Prandi et al., 2016).
Another well-established approach to the calculation
of spectral densities is the direct calculation of Huang-
Rhys factors from the gradient of the excited-state PES
at the ground-state equilibrium geometry. The peaks
of the spectral density can be broadened by gaussian or
Lorentzian functions in order to take into account the
finite vibrational lifetime and the inhomogeneous distri-
bution of vibrational frequencies (Lee et al., 2016). Usu-
ally, only the discrete part of the spectral density can
be included in such a treatment, but the effect of the
surrounding environment on the normal mode frequen-
cies and Huang-Rhys factors can be included by using
multiscale approaches (Lee et al., 2016).
At the ground-state equilibrium, the gradient of the
excited state PES is equal to the gradient of the energy
gap U . In this position, the excited-state gradient is of-
ten called vertical gradient (VG). For multiple modes,
the energy gap can be expressed, in mass-weighted coor-
dinates, as follows:
∆Ej(Q) = Ej−Eg−(cid:88)
(cid:18)
(cid:19)
ω2
ndj,nQn +
1
2
ω2
nd2
j,n
, (44)
n
where the displacement dj,n of mode n is related to
the derivative fj,n of the excited-state PES by dj,n =
−fj,n/ω2
n. The VG in normal coordinates fj can be ob-
tained from the cartesian VG, fc,j, as follows:
fj = P†M 1
2 fc,j ,
(45)
where M is a diagonal matrix containing nuclear masses
and P is a rectangular matrix whose columns are the
normal modes expressed in mass-weighted Cartesian co-
ordinates. Finally, the Huang-Rhys factor of each mode
can be calculated by
Sj,n =
λj,n
ωn
=
ωnd2
j,n
2 =
f 2
j,n
2ω3
n
.
(46)
In this explicit approach, the linear dependence of the en-
ergy gap on all normal coordinates can be assessed. More
recently, it was confirmed that the mixed approach of the
MD simulation and direct calculation of Huang-Rhys fac-
tor results in a spectral density in fairly good agreement
with experimental result (Lee and Coker, 2016). Notably,
the Huang-Rhys factors may be explicitly calculated also
for the intermolecular motions. The dimensions of the
system, however, do not allow a quantum-mechanical cal-
culation of normal modes and vertical gradients, which
26
can instead be obtained through classical modeling of the
PES and the pigment-protein interactions (Renger et al.,
2012). However, the intermolecular motions may not be
well described by harmonic potentials, and the energy
gap could be highly nonlinear with the intermolecular
coordinates.
Very recently, an extension of the method has been pre-
sented by combining VG with force-field based MD (Lee
et al., 2017): the sampled configurations are used to ini-
tiate QM ground state optimization of chromophore ge-
ometries in the presence of the instantaneous local fields
provided by the MM partial charges of the surrounding
protein environment. Ground and excited state proper-
ties are then computed at these optimized geometries to
parametrize an ensemble of instantaneous local system-
bath model Hamiltonians. The method has been success-
fully applied to two phycobiliprotein structures, namely
PE545 and PC645 complexes.
Contrary to the discrete part of the spectral density,
the continuous, low-frequency part is much more chal-
lenging to be computed at QM level, and it may require
extensive calculations along an MD trajectory. More-
over, the intermolecular motions that give rise to the
low-frequency part are strongly dependent on the tem-
perature, as some barriers become accessible only when
enough thermal energy is present in the system. There-
fore, the size and shape of the continuous spectral density
may indeed be dependent on temperature (Rancova and
Abramavicius, 2014).
FIG. 13 Bath spectral densities for BChl-3 of the FMO com-
plex. "Exp." represents the experimental ∆-FNL data in
(Ratsep and Freiberg, 2007), "Simulation (LC)" the theoreti-
cal spectral density of (Lee and Coker, 2016), and "Simulation
(KWC)" the theoretical spectral density of (Kim et al., 2018).
The inset shows close-up of the lower frequency region. All
the data have been provided by Young Min Rhee.
As an alternative approach, Rhee and coworkers re-
cently developed (Kim and Rhee, 2016; Park and Rhee,
2012) a new scheme that combines interpolated DFT-
level calculations of pigment and molecular mechanics
calculations of protein environments. This approach
constructs on-the-fly quantum portion of potential en-
ergy surface through interpolation from pre-calculated
database, and is efficient while maintaining reasonable
accuracy of the dynamics. As a result, dynamics lasting
up to about 100 ns has been shown to be possible (Kim
et al., 2018). Figure 13 compares the spectral densities
calculated in this manner with that based on the mixed
MD and HR factor calculation method (Lee and Coker,
2016). Although there is discrepancy between the two
theoretical approaches, which can be related to motional
narrowing effects and different force fields and conditions,
both of these are in reasonable agreement with the ex-
perimental results. Although the issues of anharmonic
and nonlinear coupling effects of the bath still need to be
examined more carefully, the level of agreement as can be
seen in Fig. 13 provides a strong support for the validity
of the conventional EBH for the FMO complex.
V. EXCITON-RADIATION INTERACTION, LINESHAPE
FUNCTIONS, AND RESPONSE FUNCTIONS
The total Hamiltonian for the LHC in the presence of
radiation is the sum of Eq. (20) and the matter-radiation
interaction Hamiltonian Hint(t) as follows:
HT (t) = Egg(cid:105)(cid:104)g + Hex + Hint(t) ,
(47)
where Hex is the single exciton-bath Hamiltonian defined
by Eq. (20). In general, double exciton states should also
be included for a complete description of general four-
wave mixing spectroscopy. While this is straightforward,
we omit this contribution here for the sake of simplicity.
We denote the transition dipole vector for the excitation
from g(cid:105) to sk(cid:105) as µk. Then, the total electronic polar-
ization operator for the transitions to the single exciton
space is given by
P =
µk (sk(cid:105)(cid:104)g + g(cid:105)(cid:104)sk)
(Djϕj(cid:105)(cid:104)g + g(cid:105)(cid:104)ϕjDj) ,
(48)
(cid:88)
(cid:88)
k
j
=
where Dj =(cid:80)
k µkU∗
kj.
The theories of absorption and emission lineshapes for
Eq. (47) are well established. For example, for the case
of radiation with frequency ω and polarization η, the
absorption lineshape function is given by
(cid:90) ∞
(cid:110)
(49)
j η · Djϕj(cid:105) and (cid:104)···(cid:105) represents all aver-
aging over the disorder within the ensemble of the sample
and polarization direction. For the sake of completeness,
e− i t HexD(cid:105)(cid:104)Dρbe
(cid:111)(cid:69)
I(ω) =
×(cid:68)
where D(cid:105) =(cid:80)
e
i tEg T r
dt eiωt
i t Hb
1
π
Re
0
,
27
Appendix C provides a derivation of the above expression
starting from a more general time dependent Hamilto-
nian. Alternatively, one can obtain the above expression
from the time correlation function of the polarization op-
erator, Eq. (C5). Similarly, the emission lineshape func-
tion is given by
(cid:90) ∞
(cid:110)
0
E(ω) =
×(cid:68)
1
π
Re
e− i tEg T r
dt e−iωt
e− i t HbD(cid:105)(cid:104)Dρexe
(cid:111)(cid:69)
i t Hex
,
(50)
where ρex = e−β Hex/T r{e−β Hex}. Appendix C provides
a derivation of the above expression as well.
Four wave mixing spectroscopies are useful for interro-
gating the dynamics of excitons because they offer more
selective information on the exciton space with appropri-
ate choice of pulse sequences and phase matching condi-
tions. Theories of four wave mixing and multidimensional
electronic spectroscopy are well established (Abramavi-
cius et al., 2009; Cho, 2008; Mukamel, 1995). However,
in comparison to 2D vibrational spectroscopy for which
accurate computational modeling of signals have been
shown to be feasible for a broad range of systems, accu-
rate modeling of 2DES remains challenging for most of
systems.
Appendix C provides a general expression for the third
order polarization, Eq. (C40), the major observable of
four-wave mixing spectroscopy, and detailed expressions
for its components, Eqs. (C43)-(C46). The key quanti-
ties containing the information on the material system
are the third order response functions defined by Eqs.
(C43)-(C46). 2DES is a collection of resulting third or-
der polarizations, typically represented with respect to
two frequencies of Fourier transform for the initial and
final coherence times, t2 − t1 and tm − t3, respectively
(see Appendix C for the definition of these times).
In
the absence of time dependent fluctuations and consider-
ing only transitions between the ground electronic state
and single exciton states, the four response functions can
be expressed as (see Appendix C for the details of deriva-
χ(1)(tm, t, t(cid:48), t(cid:48)(cid:48)) =
i (t(cid:48)−t(cid:48)(cid:48))Eg e− i (tm−t)Eg
e− i (tm−t(cid:48)) HbDm(cid:105)(cid:104)D2e− i (t(cid:48)−t(cid:48)(cid:48)) Hex
tion):
(cid:68)
e
×T r
× ρbD1(cid:105)(cid:104)D3e
χ(2)(tm, t, t(cid:48), t(cid:48)(cid:48)) =
(cid:68)
×T r
× ρbD1(cid:105)(cid:104)D2e
χ(3)(tm, t, t(cid:48), t(cid:48)(cid:48)) =
(cid:68)
×T r
×ρbD2(cid:105)(cid:104)D1e
χ(4)(tm, t, t(cid:48), t(cid:48)(cid:48)) =
(cid:68)
(cid:110)
(cid:110)
(cid:110)
(cid:111)(cid:69)
(cid:111)(cid:69)
(cid:111)(cid:69)
i (tm−t) Hex
i (t−t(cid:48)(cid:48)) Hb e
e− i (tm−t)Eg e
i (t(cid:48)−t(cid:48)(cid:48))Eg
,
(51)
e− i (tm−t) HbDm(cid:105)(cid:104)D3e− i (t−t(cid:48)(cid:48)) Hex
i (tm−t(cid:48)) Hex
i (t(cid:48)−t(cid:48)(cid:48)) Hb e
e− i (tm−t)Eg e− i (t(cid:48)−t(cid:48)(cid:48))Eg
e− i (tm−t) HbDm(cid:105)(cid:104)D3e− i (t−t(cid:48)) Hex
(cid:110)
i (tm−t(cid:48)(cid:48)) Hex
i (t(cid:48)−t(cid:48)(cid:48)) Hb e
e− i (tm−t)Eg e− i (t(cid:48)−t(cid:48)(cid:48))Eg
,
(52)
,
(53)
×T r
×D2(cid:105)(cid:104)D3e
e− i (tm−t(cid:48)(cid:48)) HbDm(cid:105)(cid:104)D1ρbe
i (tm−t) Hex
i (t−t(cid:48)) Hb e
i (t(cid:48)−t(cid:48)(cid:48)) Hex
(cid:111)(cid:69)
.
(54)
Figure 14 provides diagrammatic representations of the
above four response functions, which are rephasing and
non-rephasing terms of the ground state bleaching and
stimulate emission terms. Excited state absorption terms
are not shown because we limited the consideration to
only single exciton states here.
Despite recent advances, theoretical understanding of
2DES signals for the exciton states of LHCs remains chal-
lenging. Vibronic couplings of pigment molecules in pro-
tein environments are more pronounced than those of
isolated pigment molecules. Their contribution to spec-
troscopic signals can be significant even for BChls that
are known to have very small HR factors. Due to the fact
that the exciton Hamiltonian does not commute with
the exciton-bath coupling in general, a specific exciton
state ϕj(cid:105) created by an incoming photon starts decoher-
ing and relaxing to other exciton states almost immedi-
ately. These are depicted explicitly in Fig. 14 by dashed
(and blue, on line) lines. In addition, the states of bath
can be complicated in the exciton manifold because of
non-adiabatic effects associated with multiple electronic
states. In LHCs, the rates of these decoherence and re-
laxation processes, which are due to vibronic couplings,
are comparable to those due to purely electronic cou-
plings that can happen if a specific site excitation state
sj(cid:105) can be created. Furthermore, additional complica-
tions can arise due to the fact that excitations created
spectroscopically in LHCs are far from localized site ex-
citation states in general.
Even for highly tuned excitation wavelength, what is
being created is most likely to be a subensemble of exci-
ton states mixed together, which are degenerate in energy
(a)
(b)
28
(c)
(d)
FIG. 14 Diagrams representing four terms contributing to the
third order polarization. The first two terms (a) and (b) corre-
spond to echo signals satisfying the phase matching condition
of km = k3 + k2 − k1, and are in general called rephasing
terms. The last two terms satisfy a different phase matching
condition of km = k3 − k2 − k1 and are in general called non-
rephasing terms. Blue dashed curves represent relaxation and
decoherence caused by the bath that continue propagating
across interaction with radiation pulse, and Rk with different
k represents a different bath relaxation operator. Bg repre-
sents the bath of ground electronic state and Be the bath of
the manifold of the single exciton states. Different bath states
created following interaction with the radiation are expressed
as different number of primes in the superscript.
but have different extent in linear combinations of exciton
states and in vibronic contributions. Selection of specific
polarizations can reduce the size of the subensemble, but
the qualitative nature of the subensemble is expected to
remain the same. As a result, any attempt to detect
the electronic coherence directly from 2DES in time is
expected to be significantly hampered by the effect of
subensemble dephasing.
There are additional complications that can arise in the
four-wave mixing spectroscopy of excitons in LHCs. For
the simple case with a well isolated single excited state,
dephasing due to the disorder in the excitation energy is
cancelled in the rephasing signal when t2 − t1 = tm − t3.
However, as can be seen from Eqs. (51) and (52) (and
also Figs. 12(a) and (b)), the rephasing signal contains
multiple contributions involving closely spaced exciton
states. Therefore, full recovery of phase relation is not
possible in this case because the disorder affects different
exciton states within the subensemble in a different man-
ner. In addition, the contributions of vibrational modes
in both the ground electronic state and single exciton
states are non-negligible. Unlike the electronic coherence,
vibrational coherence is less susceptible to dephasing and
can survive longer.
The initial interpretation that 2DES beating signals
(Collini et al., 2010; Engel et al., 2007) might reflect
purely electronic coherence, is based on the assumption
that vibronic couplings are negligible. Even if this is the
case, such coherence may not have significant implication
in the context of exciton dynamics unless it links exci-
ton states that are spatially well separated. Let alone
the question of how much the 2DES signal reflects the
creation and evolution of excitons under natural irradia-
tion of sun light, unambiguous interpretation of beating
signals in general require clear resolution of major fea-
tures of underlying Hamiltonian governing the dynam-
ics of excitons and details of matter-radiation interaction
(Ginsberg et al., 2009). Unambiguous interpretation of
spectroscopic signals and detailed computational study
of spatio-temporal evolution of excitons are important
in this respect. Significant advances have been made in
dealing with these issues in the past decade, but much
still need to be accomplished.
VI. THEORETICAL DESCRIPTION OF EXCITON
DYNAMICS AND SPECTROSCOPIC OBSERVABLES
A. Overview of theoretical and computational approaches
Theoretical research on the dynamics of excitons has
a long history dating back to early days of quantum me-
chanics.
In particular, for excitons formed in molecu-
lar aggregates and solids, there have been extensive the-
oretical and computational efforts to understand exci-
ton migration kinetics and absorption/emission spectro-
29
scopic data (Agranovich and Galanin, 1982; Agranovich
and Hochstrasser, 1982; Kenkre and Reineker, 1982; Sil-
bey, 1976). While these have laid fundamental basis for
modern research on the dynamics of molecular excitons,
they have relied heavily on phenomenological treatment
of exciton relaxation and dephasing dynamics, and have
focused mostly on steady state properties. In addition,
because systems that were studied consisted of either
identical molecules or binary mixtures of host and guest
molecules at most, underlying exciton Hamiltonians have
had rather simple features.
With the exception of chlorosome of green sulfur bacte-
ria, which is a superaggregate of BChls, all LHCs can be
viewed as host-guest systems of 10− 100 nm length scale
sizes. Proteins serve as host environments, and pigment
molecules are guest molecules, serving as site basis for
excitations. The distinctive nature of LHCs compared to
simple molecular host-guest systems is that protein hosts
play an active role in tuning excitation energies and con-
trols spatial arrangement of pigment molecules to a great
extent. Therefore, accurate specification of the exciton
Hamiltonian, exciton-bath coupling, and bath Hamilto-
nian terms themselves are important for reliable charac-
terization of LHCs. Because the magnitudes of these pa-
rameters are in the intermediate range, well established
theoretical approaches developed for simple or limiting
situations are either inappropriate or in need of verifica-
tion by more advanced approaches. In this regard, LHCs
have motivated new advances in theories and computa-
tional methods.
Theoretical approaches to describe and simulate ex-
citon dynamics can be classified into three classes, de-
pending on the level of information and the degree of
accuracy sought after. The minimal approach is to con-
sider the time evolution of excitation/exciton popula-
tions. Broadly, this can be called master equation (ME)
approach. The next level of description is to consider
the time evolution of a reduced exciton density operator
(RDO). We here call this RDO approach. Finally, the
most complete but expensive one is to describe the time
evolution of the full density operator (FDO) represent-
ing both exciton and the bath.
In practice, this FDO
approach relies on being able to effectively represent the
bath of infinite size in terms of finite ones. Below we
provide a brief overview of the three major approaches,
and then explain how the outcomes of these theoretical
calculations enable new understanding of the nature of
excitons in LHCs.
1. Master equation approach and rate description
In ME approach, the physical observables of interest
are exciton populations, which are sufficient for calculat-
ing exciton mobility and the extent of delocalization. An
obvious choice of the unit of exciton here is each pigment.
In principle, it is possible to formally construct exact and
generalized ME (Kenkre and Knox, 1974) as follows:
{Wk→j(t)pk(t) − Wj→k(t)pj(t)} ,
(55)
(cid:88)
k(cid:54)=j
d
dt
pj(t) =
where pj(t) is the exciton population at the jth pigment
molecule and Wj→k(t) is the transfer rate of exciton pop-
ulation from the jth to the kth pigment that can be de-
fined to be formally exact by accounting for all possible
pathways. This approach can serve as a general method-
ology for all LHCs if accurate rate expressions are avail-
able. However, in practice, calculating the exact rate
kernel entails accounting for all the many body quantum
interactions and thus remains a challenging theoretical
task.
If the electronic couplings between site excitations of
pigment molecules are small compared to other parame-
ters and the steady state limit is assumed, one can ap-
proximate Wj→k(t) with the following Forster resonance
energy transfer (FRET) rate expression (Forster, 1948,
1959):
(cid:90) ∞
−∞
kF
j→k =
J 2
jk
2π2
dωLj(ω)Ik(ω) .
(56)
where we have adopted a more general definition of
FRET by assuming that Jjk is not limited to transition
dipole interactions. In the above expression, Lj(ω) and
Ik(ω) are lineshape functions for the emission of the jth
pigment and absorption of the kth pigment. For exam-
ple, let us assume that Eq. (20) can be simplified as
HLHC = Egg(cid:105)(cid:104)g +
Jjksj(cid:105)(cid:104)sk
Nc(cid:88)
(cid:16) Bjsj(cid:105)(cid:104)sj + Hb,j
j=1
Ejsj(cid:105)(cid:104)sj +
(cid:17)
Nc(cid:88)
(cid:88)
j=1
k(cid:54)=j
,
(57)
where Bj and Hb,j are exciton-bath coupling and bath
Hamiltonian localized to the site exciton state sj(cid:105). Then,
the lineshape functions introduced in Eq. (56) are defined
as
dte−iωt+i(Ej−Eg)t/
ei( Hb,j + Bj )t/
e−i Hb,j t/
e−β( Hb,j + Bj )(cid:111)
,
dt eiωt−i(Ek−Eg)t/
ei Hb,kt/
e−i( Hb,k+ Bk)t/
e−β Hb,k
(58)
,
(59)
(cid:111)
Nc(cid:88)
j=1
+
−∞
(cid:90) ∞
(cid:110)
(cid:90) ∞
(cid:110)
−∞
Lj(ω) =
× 1
Z(cid:48)
b,j
T rb,j
Ik(ω) =
× 1
Zb,k
T rb,k
b,j = T rb,j{e−β( Hb,j + Bj )} and Zb,k =
where Z(cid:48)
T rb,k{e−β Hb,k}.
(cid:90) ∞
−∞
30
The normalized emission lineshape (in the unit of ν =
ω/(2πc)) for the jth pigment is expressed as
fj(ν) = τj
25π3nj,rµ2
j c
3
ν3Lj(2πcν) ,
(60)
where nj,r is the refractive index around the jth chro-
mophore. The molar extinction coefficient for the kth
pigment is related to the lineshape function by the fol-
lowing relation:
Ik(2πcν) =
3000(ln 10)nk,r
(2π)2NAµ2
k ν
k(ν) ,
(61)
where nk,r is the refractive index around the kth pigment.
Then, Eq. (56) can be expressed as
kF
j→k =
9000(ln 10)
128π5NAτj
nj,r
nk,r
J 2
jk
j µ2
µ2
k
dν
fj(ν)k(ν)
ν4
. (62)
This expression becomes the well-known Forster's spec-
tral overlap expression (Forster, 1948) in the limit where
Jjk is due to the transition dipole-dipole interaction
and the dielectric constants around jth and kth chro-
mophores are the same.
More general expressions than Eq.
(56), which in-
clude nonequilibrium (Jang and Cheng, 2013; Jang et al.,
2002b) and inelastic effects (Jang, 2007; Jang and Cheng,
2013), are also available. However, the FRET rate ex-
pression or its generalizations, which are based on the
second order approximation with respect to the electronic
coupling between a pair of pigment molecules, are inap-
propriate for many LHCs because not many excitons are
localized at single chromophores.
Alternatively, one can extend the unit of exciton pop-
ulation as that residing in a group of strongly coupled
pigment molecules.
If groups of chromophores can be
identified such that electronic couplings between pigment
molecules in different groups are weak enough, the rates
of exciton transfer between them can be calculated, again
employing a second order approximation with respect to
the inter-group electronic couplings. This is the idea be-
hind the multichromophoric FRET (MC-FRET) (Jang
et al., 2004; Sumi, 1999), which was recently re-derived
and tested more extensively (Ma and Cao, 2015).
The use of MC-FRET rates as kernels of ME was for-
mulated more rigorously by introducing the concept of
modular excitons (Jang et al., 2014). Under the assump-
tion that all the pigment molecules constituting an LHC
can be divided into disjoint modules and that only the
coarse-grained overall exciton population of each mod-
ule, pn(t), (modular exciton density) is of interest, one
can consider the following generalized ME:
(cid:110) Wm→n(t)pm(t) − Wn→m(t)pn(t)
(cid:111)
(cid:88)
.
d
dt
pn(t) =
m(cid:54)=n
(63)
As in the case of Eq. (55), exact formal expression for
Wn→m(t) can be found easily (Jang et al., 2014). In prac-
tice, approximations are needed. Under the assumption
that the inter-module electronic couplings are small com-
pared to other parameters and that the exciton density
within each module reaches the stationary limit quickly,
it is possible to show that (Jang et al., 2014)
Wn→m(t) =
Jj(cid:48)k(cid:48)Jj(cid:48)(cid:48)k(cid:48)(cid:48)
2π2
dωLn
j(cid:48)(cid:48)j(cid:48)(t, ω)I m
k(cid:48)k(cid:48)(cid:48) (ω) ,
0
formally exact QME is well-known:
31
σI (t) =
(cid:26)
d
dt
−iT rb
(cid:90) t
−
(cid:20)
−i
(cid:90) t
(cid:20)
0
(cid:26)
Leb,I (t) exp(+)
(cid:27)
Leb,I (t) exp(+)
×Q Leb,I (τ )ρb
dτ T rb
(cid:21)
(cid:27)
(cid:21)
dτQ Leb,I (τ )
QρI (0)
−i
dτ(cid:48)Q Leb,I (τ(cid:48))
(cid:90) t
τ
σI (τ ) .
(68)
(cid:88)
(cid:88)
j(cid:48)j(cid:48)(cid:48)
k(cid:48)k(cid:48)(cid:48)
(cid:90) ∞
−∞
j(cid:48)(cid:48)j(cid:48)(t; ω) and I m
(64)
where Ln
k(cid:48)k(cid:48)(cid:48)(ω) are lineshape matrix ele-
ments representing the time dependent emission of mod-
ule n and the absorption of module m. These are respec-
tively expressed as
(cid:20)(cid:90) t
(cid:110)(cid:104)sn,j(cid:48)(cid:48)e−i Hb,nt(cid:48)/
j(cid:48)(cid:48)j(cid:48)(t; ω) ≡ 2Re
Ln
(cid:90) ∞
× T rbn
(cid:110)(cid:104)sm,k(cid:48)ei Hb,mt(cid:48)/
k(cid:48)k(cid:48)(cid:48)(ω) ≡
I m
×T rbm
−∞
0
dteiωt+iEgt/
dt(cid:48)e−iωt(cid:48)−iEgt(cid:48)/
ρnsei Hnt(cid:48)/sn,j(cid:48)(cid:105)(cid:111)(cid:105)
bmsm,k(cid:48)(cid:48)(cid:105)(cid:111)
e−i Hmt(cid:48)/
ρg
, (65)
,(66)
where Hn ( Hm) and Hb,n ( Hb,m) are respectively the
EBH in the single exciton space and the bath Hamilto-
nian of the nth (mth) module. In the limit of t → ∞, Eq.
(64) approaches the MC-FRET rate (Jang et al., 2014,
2004).
2. Reduced density operator approach
For pigment molecules with moderate or strong elec-
tronic couplings, description of the dynamics at the level
of a reduced density operator (RDO) in the exciton man-
ifold serves as a better platform than restricting the focus
on the exciton population only. Let us denote the RDO
at time as σ(t). Then, the quantum master equation
(QME) governing the time evolution equation of σ(t) can
in general be expressed as
d
dt
σ(t) = − i
[ He, σ(t)] − R[t; σ] + I[t] ,
(67)
where the first term represents the dynamics due to the
exciton Hamiltonian, R[t; σ] accounts for the relaxation
and dephasing due to environments, and I[t] is the in-
homogeneous term that reflects the effect of the initial
condition.
There is a large body of literature available on the
derivation of formally exact QME and approximations.
For example, for the case of the interaction picture RDO
defined as σI (t) = T rb{ρI (t)}, where ρI (t) is the total
density operator in the interaction picture, the following
(cid:90) t
(cid:90) t
0
d
dt
−
−
0
In the above equation, Leb,I (·) = [ Heb,I (t), (·)]/, Q(·) =
1 − T rb{(·)}.
Equation (68), although exact, is not useful in prac-
tice because exp(+)[−i(cid:82) t
τ dτ(cid:48)Q Leb,I (τ(cid:48))] cannot be deter-
mined without full information on the system and bath
degrees of freedom. For practical calculations, approxi-
mations are made in general, although recent theoretical
advances (Cohen et al., 2013; Kelly et al., 2016; Shi and
Geva, 2004; Zhang et al., 2006) suggest that exact nu-
merical evaluation of this term is feasible.
One of the simplest and most popular approximation
is the 2nd order approximation with respect to Leb,I (t),
which results in
(cid:110) Lsb,I (t)ρI (0)
(cid:111)
(cid:110) Lsb,I (t) Lsb,I (τ )QρI (0)
(cid:111)
σI (t) = −iT rb
dτ T rb
dτ T rb {Lsb,I (t)Lsb,I (τ )ρb} σI (t∗) ,
(69)
where t∗ can be either τ or t.
In the former case, the
above equation becomes time-nonlocal (TN). In the lat-
ter case, it becomes time-local (TL). Which choice works
better depends on the nature of the system-bath cou-
plings and the bath dynamics, but most numerical exam-
ples so far suggest that the TL equation performs bet-
ter than the TN equation, at least at room temperature
where the bath dynamics become most likely Gaussian
(Chen et al., 2009; Palenberg et al., 2001). Analyses of
the fourth order QME (Jang et al., 2002a) and steady
state limits (Fleming and Cummings, 2011) suggest that
this is because the 2nd order TL equation can account
for some of the effects of the 4th order terms of TN equa-
tions.
In Eq. (69), the super-operator involving the second
order correlations of Lsb(t), which appears commonly in
both inhomogeneous and homogeneous terms, can be ex-
pressed as
(cid:90) t
(cid:88)
0
=
dτ T rb
(cid:88)
(cid:110) Lsb,I (t) Lsb,I (τ )ρb
(cid:90) t
(cid:111)
(·)
dτCjk(t − τ )[sj(t), sk(τ )(·)] + H.c. ,
j
k
0
(70)
where H.c. denotes Hermitian conjugates of all the pre-
vious terms, sj(t) = eiHet/sj(cid:105)(cid:104)sje−iHet/
, and Cjk(t) is
the bath correlation function for sites j and k defined as
Cjk(t) =
ngj,rgk,r
(cid:88)
(cid:90) ∞
r
(cid:8)(bre−iωt + b†
(cid:18)
ω2
×T rb
dωJjk(ω)
coth(
(cid:9)
(cid:19)
r)ρb
reiωt)(br + b†
βω
2
) cos(ωt) − i sin(ωt)
,
=
1
π
0
(71)
where Jjk(ω) is the bath spectral densities defined by Eq.
(22). The 2nd order QMEs in various forms and approx-
imations have served as major theoretical tools for de-
scribing exciton dynamics and calculating spectroscopic
data in early pioneering quantum dynamical studies on
LHCs, and still remain as key theoretical methods offer-
ing qualitative and/or semiquantitative information.
As is well-known, the 2nd order approximation with
respect to the system-bath interaction causes the QME
to be non-positive definite, without rotating wave ap-
proximation (Kohen et al., 1997; Pechukas, 1994), and
to be unreliable as the system-bath interaction becomes
larger. An alternative approach addressing this issue but
still based on the assumption of weak system-bath cou-
pling is the Lindblad equation (Lindblad, 1976). Having
been constructed axiomatically, the Lindblad equation
guarantees complete positivity and thus can be used to
model the effects of environments on any quantum sys-
tem. For this reason, it has been used extensively in
the quantum information community and was brought
to the study of the FMO complex following the 2DES
spectroscopy (Engel et al., 2007). However, due to the
assumption of Markovian-bath intrinsic in the theory and
the phenomenological nature of relaxation and dephas-
ing super-operators used, it has primarily served as a
useful analysis tool rather than a quantitative modeling
method. Of course, it is possible to obtain some micro-
scopic expressions for the terms of Lindblad equation,
for example, from the 2nd order QME. However, this
involves additional approximations the physical implica-
tions of which are not always clear.
Considering typically moderate nature of system-bath
interactions found in most LHCs, it is not likely that the
errors caused by QMEs based on perturbative approxi-
mations are substantial. However, in general, assessment
of the accuracy of the 2nd order QMEs is possible only
when benchmarked against more accurate approaches.
For this reason, different approaches employing higher
order approximations or nonperturbative methods have
been applied to LHCs.
Among various higher order methods developed
for decades,
the hierarchical equations of motion
(HEOM) approach (Tanimura, 2006, 2015; Tanimura and
Mukamel, 1993) has gained popularity recently and its
results are considered as benchmark data by many re-
32
searchers. The HEOM approach was originally formu-
lated and developed within the path integral influence
functional formalism for open system quantum dynam-
ics, and is based on the idea that higher order terms
of the system-bath interactions can be accounted for
by introducing a hierarchy of coupled auxiliary opera-
tors, an assumption valid as long as the bath correlation
functions have exponential functional form. With fur-
ther advances in practical schemes (Ishizaki and Flem-
ing, 2009b; Schroter et al., 2015; Tanimura, 2006; Zheng
et al., 2012) to bring closure to the hierarchy and com-
putational power, application of the HEOM approach to
LHCs became feasible recently.
There are also other approaches that have been de-
veloped more recently and can account for higher order
system-bath interactions in different manner. These in-
clude polaron-transformed QME approach (Jang, 2011;
Jang et al., 2013; McCutcheon and Nazir, 2011; Nazir,
2009), generalized QME approach (Cohen et al., 2013;
Kelly et al., 2016; Shi and Geva, 2004; Zhang et al.,
2006), and transfer tensor approach (Kananeka et al.,
2016; Rosenbach et al., 2016). Well-known variations
of QME approaches with modified definitions of system
(Iles-Smith et al., 2016, 2014) or system-bath couplings
(Hwang-Fu et al., 2015; Novoderezhkin and van Gron-
delle, 2017) have also been demonstrated to be accurate
enough with appropriate corrections.
While QME approaches are in general appropriate
for describing the effects of quantum mechanical bath,
they tend to neglect the effects of classical and stochas-
tic fluctuations, which cannot be captured well in the
form of Hamiltonian but may still be substantial in ac-
tual environments. For example, for LHCs embedded
in membrane, there can be various noises being prop-
agated from distant sources but having effects on dy-
namics around pigment molecules. In addition, some of
anharmonic and nonlinear effects of the bath, which are
neglected in constructing the model Hamiltonian with
harmonic oscillator bath, may still affect the dynamics
by appearing in the form of uncontrollable noises due
to the chaotic/irregular nature of the trajectories they
incur in general multidimensional space. The effects of
these noises can be significant at ambient condition and
may have to be included along with the quantum bath.
While the contributions of these noises to the exciton dy-
namics can be treated at a simple phenomenological level
such as Haken-Strobl equation (Haken and Strobl, 1973),
more satisfactory QME level description incorporating
them into a consistent quantum description of bath is
not available to the best of our knowledge. Functional
integration (Ritschel et al., 2011) and Heisenberg picture
time evolution (Ghosh et al., 2009) approaches, which in-
voke somewhat different approximations but are general
otherwise, may be better suited to this end.
3. Full density operator approaches
The QMEs and the HEOM approaches, despite having
been successful, are limited mostly to the harmonic oscil-
lator bath model. Little is understood regarding the ef-
fects of anharmonic contribution of the bath Hamiltonian
on the nature and dynamics of excitons in LHCs. For sat-
isfactory account of these effects, it is often necessary to
consider the dynamics of the full density operator (FDO)
representing the system and the bath. In addition, even
for the bath of harmonic oscillators, accurate calculation
of higher order response functions in general require time
evolution of FDO unless a new kinds of coupled QMEs
are developed. Well-known FDO approaches are semi-
classical dynamics methods (Cotton and Miller, 2016;
Huo and Coker, 2010; Miller, 2001), mixed quantum-
classical approaches (Tully, 2012; Zheng et al., 2017) such
as surface hopping and mean field approximations. While
these offer more realistic treatment of the bath degrees of
freedom and their interactions with excitons, they entail
different levels of approximations in the quantum dynam-
ics. In addition, the bath has to be finite for practical
calculations. Given that these approximations and issues
are well addressed, the results of these FDO approaches
can offer important information for LHCs that are not
accessible through QME and HEOM approaches.
4. Computational methods for lineshape and response functions
All the QME approaches can be used to calculate the
absorption and emission lineshapes, employing the prod-
uct of the transition dipole and the density operator in
either ground or excited state as an initial condition. Al-
ternatively, a specific QME method can be modified to
derive an appropriate lineshape theory using the coher-
ent term of the density operator as the initial condition
and applying mixed time evolutions, one on the ground
electronic state and the other on the manifold of excited
electronic states, respectively, to the two different sides of
the density operator. One popular approach for calculat-
ing the lineshape is the so called modified Redfield equa-
tion (Kleinekathofer and Schreiber, 2006; Zhang et al.,
1998), which includes the diagonal components of the
exciton-bath coupling as part of the effective system part
and treat the off-diagonal components in a perturbative
manner. Most recently, this approach was extended fur-
ther to account for the finite relaxation time of nuclei
(Dinh and Renger, 2016). Modified Redfield equation
can also be extended for the simulation of exciton dy-
namics (Hwang-Fu et al., 2015; Novoderezhkin and van
Grondelle, 2017; Yang and Fleming, 2002), although care
should be taken in identifying the unit of population in
this case.
An important issue is that response functions of four-
wave mixing spectroscopy cannot be calculated exactly
33
even at a formal level using a conventional QME ap-
proach without modification. The diagrams in Fig. 14
clarify the reason for this. Each interaction with pulse
corresponds to the creation of a non-equilibrium state,
which also depends on previous time evolution of both
excitons and the bath that are again dependent on the
nature of previous pulses. For this reason, the relaxation
and dephasing terms due to the bath are expected to be
functionals of whole sequence of pulses, which are in gen-
eral different from those of equilibrium bath. Therefore
an FDO approach is necessary in general (McRobbie and
Geva, 2009). A perturbative approach addressing this
issue has also been formulated (Jang, 2012).
B. Applications
1. FMO complex
Early theoretical works on the FMO complex employed
the 2nd order ME and QME approaches. Application
of the Redfield equation (Redfield, 1957), the 2nd order
TL-QME in the Markovian bath limit, with the secu-
lar approximation (Renger and May, 1998) showed that
the approach can reproduce the major features of tem-
perature dependence in the absorption lineshape and
the time dependence of pump-probe spectroscopic data.
They also provided estimate for the relaxation of a higher
exciton state (4th state) to be about 2 ps. This ap-
proach was later combined with an optimal control the-
ory (Bruggemann and May, 2004) to explore the possi-
bility of controlling the exciton dynamics in the FMO
complex through pulse shaping. On the other hand, the
ME approach in the exciton basis was employed (Vulto
et al., 1999) to describe time resolved absorption differ-
ence spectra and the exciton population dynamics. They
have also identified various relaxation time scales of exci-
ton dynamics ranging from sub 100 fs up to about 2.3 ps.
The phenomenological assumption used in this model
Hamiltonian was later justified microscopically (Adolphs
et al., 2008).
The suggestion of quantum information processing
(Engel et al., 2007) motivated new theoretical works
(Caruso et al., 2009, 2010; Chin et al., 2010; Mohseni
et al., 2008; Palmieri et al., 2009; Pelzer et al., 2012; Ple-
nio and Huelga, 2008; Rebentrost et al., 2009b,c; Sarovar
et al., 2010; Skochdopole and Mazziotti, 2011) aimed at
understanding the role of entanglement, quantum coher-
ence, and noise. Earlier versions of these (Caruso et al.,
2009; Mohseni et al., 2008; Plenio and Huelga, 2008;
Rebentrost et al., 2009c) employed the Lindblad equation
(Lindblad, 1976) or its stochastic implementation, which
are based on the assumption of weak system-bath cou-
pling and Markovian bath. Later works improved these
by including the effects of non-Markovian bath (Caruso
et al., 2010; Chin et al., 2010; Rebentrost et al., 2009a),
correlated bath fluctuations (Rebentrost et al., 2009b),
and full consideration of system-bath coupling at the
level of HEOM (Sarovar et al., 2010). According to these
studies, the effect of entanglement, which is limited to
mode entanglement, is not significant for single excitons
in the FMO complex. On the other hand, dephasing and
noise were identified as significant factors for the overall
efficiency of energy transport, as espoused by new terms
such as environment-assisted quantum transport (Lam-
bert et al., 2013; Rebentrost et al., 2009c) and dephasing-
assisted transport (Chin et al., 2010). Studies based on
the Haken-Strobl equation (Hoyer et al., 2010; Vlaming
and Silbey, 2012) have also offered new insights into the
potential complications and effects of noise. It was also
suggested that the redundancy (Skochdopole and Mazz-
iotti, 2011) of energy transfer pathways plays an impor-
tant role.
Due to the intermediate nature of the exciton-bath
coupling in the FMO complex, the Lindblad equation and
the 2nd order ME/QME approaches were not perceived
as quantitatively reliable. HEOM calculations were per-
formed to investigate the population dynamics (Ishizaki
and Fleming, 2009a), which served as the first set of
key benchmark data in this respect. This work showed
that coherent population dynamics persists up to sev-
eral hundred femto seconds even at room temperature.
Calculations (Chin et al., 2013) based on density matrix
renormalization group technique (Prior et al., 2010) also
support this result and suggest the contribution of sig-
nificant nonequilibrium effects. Later, other approaches
have been employed to calculate the excitation dynam-
ics. For example, non-Markovian quantum state diffusion
(Ritschel et al., 2011), mixed quantum-classical Poisson
bracket mapping equation approach (Kelly and Rhee,
2011), and quasiclassical dynamics (Cotton and Miller,
2016) have been shown to produce results in reasonable
agreement with the HEOM approach. These approaches
are applicable to more general exciton-bath couplings and
bath Hamiltonians, and can be used to understand the
effects of anharmonic effects of the bath modes and non-
linear exciton-bath couplings.
Theoretical observation of coherent real time dynam-
ics for the FMO complex was an important step for elu-
cidating the quantum dynamical details of exciton dy-
namics. However, such oscillatory population dynamics
were mostly between proximate BChls with significant
electronic couplings, and did not yet imply long range
"wave-like" motion of excitons. On the other hand, for
the overall description of exciton dynamics and calcula-
tion of ensemble-averaged spectroscopic observables, sim-
pler perturbative ME and QME approaches seemed to
offer reasonable answers (Moix et al., 2011; Schmidt am
Busch et al., 2011; Wu et al., 2012). Most recently, a
full HEOM calculation for trimers (Wilkins and Dattani,
2015) demonstrated that the time scale of exciton trans-
port is similar to that based on a simple ME approach us-
34
ing FRET rate as its kernel. This result suggests that the
interpretation based on earlier theoretical studies based
on approximate theories is valid in general.
There have been further progress in theoretical model-
ing of spectroscopic data. While theoretical calculation of
linear spectroscopic data was well established and could
be used for the refinement of model parameters (Adolphs
and Renger, 2006; Jansen and Knoester, 2009; Schmidt
am Busch et al., 2011), accurate modeling and unambigu-
ous interpretation of 2DES signals has remained challeng-
ing. One of the earliest simulations of 2DES was based
on equation of motion approach, while assuming simple
Redfield tensor and secular approximation for the bath
relaxation (Sharp et al., 2010). This suggested that it
is difficult to explain the off-diagonal beating signal ob-
served from 2DES (Engel et al., 2007). A different the-
oretical modeling based on the secular Redfield equation
led to a similar conclusion, while leaving the possibil-
ity of contribution of correlated bath (Abramavicius and
Mukamel, 2011). Another theoretical modeling based on
the Redfield equation and including the vibronic states
explicitly suggested that the beating signal may have
originated from vibrational motion (Christensson et al.,
2012). HEOM calculations of absorption spectra (Hein
et al., 2012) and 2DES echo spectra (Hein et al., 2012;
Kreisbeck and Kramer, 2012) on the other hand provided
support for coherence signal of electronic nature lasting
up to about 500 fs, but in the absence of static disor-
der and other sources of fluctuations. Although based on
rather simple model Hamiltonian, careful analysis (Ti-
wari et al., 2013) demonstrated that electronic coherence
can easily dephase when the effect of reasonable magni-
tude of disorder is taken into consideration, whereas beat-
ing that originates from vibrational motion can last much
longer. Th contribution of vibronic terms can be an-
other possibility (Mourokh and Nori, 2015; Plenio et al.,
2013). As yet, because of the sensitivity of 2DES signals
to various factors and the lack of fully reliable model and
quantum dynamics methods, a definite assessment of the
implications of 2DES signals remains open.
From the functionality point of view, to what extent
the quantum effects contribute to the overall efficiency
of the exciton migration remains a central question even
to date (Wu et al., 2012). Accurate determination of the
spectral density of the bath has a significant implication
in answering this question. High frequency modes of the
spectral density in general do not affect the mechanis-
tic details of exciton dynamics (Abramavicius and Abra-
mavicius, 2014). However, low frequency modes with en-
ergies comparable to the exciton band width and their
coupling strengths can be detrimental to the exciton dy-
namics mechanism and the interpretation of 2DES spec-
troscopic data. There have been various efforts to address
these issues through all-atomistic simulations (Olbrich
et al., 2011a; Shim et al., 2012) and direct calculation of
Huang-Rhys factors (Adolphs and Renger, 2006). The
contributions of correlated fluctuations (Olbrich et al.,
2011a) and non-Gaussian bath fluctuations (Jansen and
Knoester, 2009) were also suggested as potentially sig-
nificant factors that can complicate the exciton dynam-
ics. As yet, depending on the approximations and as-
sumptions involved, the assessment of the bath spectral
density can be different.
It was once suggested that
the exciton-bath coupling might be too strong to allow
any coherent dynamics unless the dynamics starts from
a coherent initial state (Muhlbacher and Kleinekathofer,
2012). However, more recent calculations based on differ-
ent methods indicate weaker bath spectral densities that
are more in tune with older models. Ultimately, on-the-
fly ab initio nonadiabatic dynamics but with sufficient
accuracy may be necessary to settle this issue. New de-
velopment in first principles linear scaling ab initio calcu-
lation can be a promising avenue to explore in achieving
this goal (Cole et al., 2013).
2. LH2 complex
Soon after the structural information of the LH2 com-
plex became available, advanced quantum dynamical
theories, for example, addressing the exciton coherence
length (Leegwater, 1996; Meier et al., 1997c), super-
radiance (Meier et al., 1997c), and four-wave mixing
spectroscopy signals (Meier et al., 1997b; Zhang et al.,
1998) have been developed. Although based on exciton-
bath Hamiltonians of appropriate features but yet with
quite simplified forms, these works have not produced
quantitative modeling of linear spectroscopic data.
In-
stead, earlier attempts to fit experimental
lineshapes
used simple exciton Hamiltonians with disorder terms
(Hu et al., 1997), while dressing each exciton peak with
phenomenologically chosen lineshape functions (Alden
et al., 1997; Georgakopoulou et al., 2002; Wu and Small,
1998). These produced reasonable fitting of linear ensem-
ble lineshapes, which indicates that the inhomogeneous
broadening due to disorder is large enough to screen the
line broadening due to the exciton-bath coupling. As
yet, the disorder is still in the moderate regime that
makes it difficult to tell what types of disorder, diago-
nal or off-diagonal in site excitation basis, are dominant
(Jang et al., 2001). In addition, detailed information on
the disorder was shown to be important for proper in-
terpretation of nonlinear spectroscopic data (Yang et al.,
2001).
The SMS data (van Oijen et al., 1999) for the LH2
complex offered the first direct evidence for the delocal-
ized nature of excitons in its B850 unit, and offered new
opportunities and challenges for theoretical and compu-
tational modeling. On one hand, large gaps between two
major excitonic peaks of the B850 unit were inexplica-
ble based on existing exciton models. This initially led
to the suggestion of elliptic distortion of LH2 (Demp-
35
ster et al., 2001; Mostovoy and Knoester, 2000; van Oi-
jen et al., 1999). Later, it was explained by models with
modulation in site energies (Hofmann et al., 2004; Kete-
laars et al., 2001) and correlated disorder (Jang et al.,
2011). Another important theoretical issue was eluci-
dating details of SMS lineshapes. The first step to un-
derstand the physical implication of such lineshapes is
to calculate the effects of exciton-bath coupling at each
single LH2 level. To this end, an exciton-bath model
(Jang and Silbey, 2003a) incorporating the bath spectral
density (Renger and Marcus, 2002) extracted from spec-
troscopic data was developed. Lineshapes were then cal-
culated by using the 2nd order TN-QME approximation
(Jang and Silbey, 2003a,b). The features of the theoreti-
cal line shapes (Jang and Silbey, 2003a) were in qualita-
tive agreement with experimental ones (van Oijen et al.,
1999). However, the widths of the former were much
narrower than the latter. This was attributed to fluc-
tuations during the long SMS measurement time scales,
which are possible even in the very low temperature limit
because of numerous excitation and de-excitation pro-
cesses needed to collect the photons. Another source for
the discrepancy is the approximate nature of the 2nd
order TN-QME. Indeed, later calculations of absorption
lineshapes by alternative or higher order methods (Chen
et al., 2009; Kleinekathofer and Schreiber, 2006), while
based on somewhat different bath spectral densities and
calculated at higher temperatures, showed that the 2nd
order TN-QME approach underestimates the broadening
due to exciton-phonon couplings. Application of these
approaches with improved spectral densities may allow
better explanation of SMS lineshapes.
In parallel, there have been advances in the modeling of
2DES spectroscopy data of the LH2 complex, for exam-
ple, providing intriguing insights into anharmonic effects
of the bath (Rancova and Abramavicius, 2014). More re-
cently, all-atomistic modeling has offered detailed infor-
mation on physical implications of 2DES signals (Segatta
et al., 2017; van der Vegte et al., 2015). The dark states
of carotenoids of LH2, which had long been speculated
to be present, was investigated recently based on quan-
tum chemical calculations and simulation of 2DES spec-
troscopic data (Feng et al., 2017; Segatta et al., 2017).
However, the results obtained so far are not yet sufficient
to lead to a general consensus. Thus, further investiga-
tions are needed.
In a given LH2 complex, the exciton dynamics within
B800, between B800 and B850, and within B850 have
all different characteristics. It is generally accepted that
the dynamics within B800 can be well described by a
hopping model of excitons localized at each BChl. The
exciton transfer from B800 to B850 can also be described
by hopping dynamics, but the coherent delocalization of
the exciton in the B850 unit has to be taken into con-
sideration. Two early theoretical studies (Mukai et al.,
1999; Scholes and Fleming, 2000) accounted for this by
employing approximate versions of the MC-FRET theory
(Jang et al., 2004; Sumi, 1999), which included only the
contribution of population terms in the exciton basis and
used phenomenological lineshape functions. Later works
(Jang et al., 2004, 2007) employing a more satisfactory
exciton-bath Hamiltonian (Jang and Silbey, 2003a) and
full MC effect within the 2nd order TN-QME approach
(Jang and Silbey, 2003a,b) confirmed that neglecting co-
herence terms in the exciton basis of the MC-FRET the-
ory does not have significant effect on the distribution of
transfer rates. In addition, these works (Jang et al., 2004,
2007) provided more solid evidence that the theoretical
results based on the exciton-bath model (Jang and Sil-
bey, 2003a) and the MC-FRET theory (Jang et al., 2004;
Sumi, 1999) are indeed consistent with experimental re-
sults (Jimenez et al., 1996; Pullerits et al., 1997).
The exciton dynamics within the B850 unit requires
full quantum dynamical approach. Although hoping dy-
namics (Abramavicius et al., 2004) and Haken-Strobl ap-
proach (Liuolia et al., 1997) provided some insights, ac-
curate theoretical description has remained challenging.
Recent applications of the HEOM approach (Chen et al.,
2009; J. Strumpfer and K. Schulten, 2012; Strumpfer and
Schulten, 2009; Yeh et al., 2012) are important advances
in this respect, and confirm fast decoherence and relax-
ation of excitons in the range of 100 − 200 fs time scales
as seen experimentally (Agarwal et al., 2002; Book et al.,
2000). However, these calculations are still based on
rather simplified spectral densities.
In addition, these
have made no or limited consideration of the effects of
the disorder. On the other hand, it was demonstrated
that the ensemble dephasing due to disorder, while using
relatively simple dynamics theory, is sufficient to explain
fast anisotropy decay in LH2 (Stross et al., 2016). Thus,
to clarify which of the two factors determines major time
resolved experimental data on the B850 unit, large scale
HEOM calculations sampled over sufficient number of re-
alizations of the disorder is necessary. Some advances
have already been made in this direction for the calcu-
lation of absorption and emission lineshapes (Jing et al.,
2013).
The dynamics of excitons between LH2s and aggre-
gates of LH2s have important implications for the overall
energy collection efficiency, and has been subject to var-
ious computational studies (Caycedo-Soler et al., 2010;
J. Strumpfer and K. Schulten, 2012; Jang et al., 2014,
2015; Ritz et al., 2001; Strumpfer and Schulten, 2009;
Xiong et al., 2012; Yang et al., 2012). The earliest in this
endeavor was the kinetic Monte Carlo simulation of exci-
ton dynamics in simple aggregates of LH2, LH3, and LH1
(Ritz et al., 2001). In this work, a fixed value of LH2-
LH2 exciton transfer time of 10 ps was used, which was
calculated based on an approximate version of the MC-
FRET theory (Jang et al., 2004; Sumi, 1999) for a single
LH2-LH2 distance (Ritz et al., 2001). This transfer time
was also adopted in later simulation of exciton dynam-
36
ics in the entire PSU of purple bacteria (Caycedo-Soler
et al., 2010). While the value has been validated to be in
reasonable agreement with an accurate calculation based
on the HEOM approach (Strumpfer and Schulten, 2009),
it is not yet representative of all the inter-LH2 exciton
transfer dynamics that are plausible. An HEOM cal-
culation demonstrated significant potential effect of the
correlation of the bath (J. Strumpfer and K. Schulten,
2012). Application of the GME-MED approach (Jang
et al., 2014, 2015) showed that the rate can change signif-
icantly depending on the realization of the disorder and
the inter-LH2 distances that are consistent with AFM im-
ages. Figure 15 shows the distribution of rates at 300 K
between two LH2 complexes separated by a center-to-
center distance of 7.5 nm. In this result, the distribution
of rates is purely due to the Gaussian disorder in the ex-
citation energy of each BChl with standard deviation of
200 cm−1. Consideration of the distribution of inter-LH2
distances will result in wider distribution of rates (Jang
et al., 2015).
FIG. 15 Distribution of LH2-LH2 rates (left panel) and av-
erage population decay (right panel) of exciton on the initial
LH2. More details can be found in (Jang et al., 2015).
3. PBPs
Redfield
equation
Modified
employed
was
linear spec-
(Novoderezhkin et al., 2010) to model
troscopic data,
excitation anisotropy measurement,
and transient absorption spectra, and then to deduce
excitation energy transfer kinetics based on these
modelings. Following the observation of the beating
signal
from the 2DES result (Collini et al., 2010),
an FDO quantum dynamics method called iterative
density matrix propagation approach was used for the
simulation of the exciton population dynamics (Huo and
Coker, 2011). The results of this work supported the
observation from the 2DES (Collini et al., 2010) that the
electronic coherence between central dimers lasts up to
about 400 fs, but also showed that such coherence does
not have significant effect on the exciton population
dynamics of peripheral pigment molecules. On the other
hand, it was also suggested that quantum vibrational
modes have significant effect on the exciton transfer
(O'Reilly and Olya-Castro, 2014).
QM/MM approaches (Aghtar et al., 2017, 2014; Cu-
rutchet et al., 2013; Lee et al., 2017) were used to model
spectroscopic data and to simulate the exciton transfer
dynamics. The earliest effort was focused on investigat-
ing the nature of the bath spectral density, and employed
a quantum-classical dynamics simulation by conducting
wave-packet dynamics on fluctuating potential energy
surfaces (Curutchet et al., 2013). It was also confirmed
that significant portion of the bath spectral density re-
sults from the intramolecular vibrations (Aghtar et al.,
2014). The population dynamics were also shown to be
non-oscillatory, even much less than that of the FMO
complex.
The methodology of QM/MM simulation has pro-
gressed to the level of constructing comprehensive EBH
models for PBPs of both PE545 and PC645 with rea-
sonable accuracy (Lee et al., 2017). These models were
then combined with lineshape theories employing the 2nd
order TL-QME approach so as to model the absorption
and CD spectra (Lee et al., 2017), demonstrating satis-
factory agreement between theory and experiment. Anal-
yses based on these calculations confirmed significant ef-
fects of vibronic couplings, and also suggested potential
importance of correlated fluctuations of the site excita-
tion energies. A different QM/MM study, also including
the effects of the polarizable medium, showed that fluc-
tuations in couplings and site energies do not have sig-
nificant effects on the overall exciton dynamics (Aghtar
et al., 2017). Theoretical analysis of a recent 2DES spec-
troscopy for PC645 (Dean et al., 2016) also suggests that
vibronic redistribution enhances the exciton transfer rate
by about a factor of 3.5.
VII. DISCUSSION
LHCs have characteristics and challenges that are dis-
tinctively different from any other biological complexes
in relation to their specific function. While the impor-
tance of structural information for LHCs is inarguable, it
is necessary to recognize that the functionality of an LHC
is determined by how and where excitons are formed and
how fast and efficiently the excitons move from one re-
gion to another, while conserving their energies as much
as possible even in the presence of disorder and fluctua-
tions. Elucidating these details require reliable informa-
tion on the energetics and the dynamics of excitons in the
excited potential energy surfaces. In addition to struc-
tural data, a comprehensive collection of spectroscopic
and computational data are needed even for establishing
37
a Hamiltonian that can represent the behavior of excitons
with reasonable accuracy. Calculation and simulation of
quantum dynamical evolution of excitons in these sys-
tems are equally challenging due to their complexity and
sizes. Furthermore, the disorder and environmental fluc-
tuations are important factors to consider because they
can have significant effects on how excitons are formed
and evolve in time. They also make significant contribu-
tions to spectroscopic observables.
Ideally,
a spectroscopic measurement with both
nanoscale spatial resolution and femtosecond time res-
olution, or an on-the-fly ab initio quantum dynamical
simulation with sufficient accuracy (with errors less than
100 cm−1 in energy) and efficiency (to run up to tens of
picosecond and to be repeated over many realizations
of disorder) is needed to probe excitons in LHCs di-
rectly. However, in the absence of feasible spectroscopic
and computational tools meeting these needs at present,
the best strategy is to develop a well-designed theory-
experiment collaboration focused on addressing promi-
nent issues and testing important assumptions. Advances
made for the FMO complex, the LH2 complex, and the
PBPs of cryptophyte algae as reviewed here, provide
lessons and future guidelines that can be applied to other
LHCs with more complexity and larger scales.
The results of various investigations on these three
LHCs serve as concrete examples for addressing many im-
portant issues concerning the excitons in LHCs. Among
them, we discuss here three topics that are in fact in-
timately connected together and have generated heated
discussion and investigation in the past decade, namely,
optimality (Cao and Silbey, 2009; Jang and Cheng, 2013;
Jang et al., 2007, 2015; Mohseni et al., 2014), quantum
coherence (Kassal et al., 2013; Olaya-Castro et al., 2012)
and robustness (Chen et al., 2013; Cleary et al., 2013;
Huh et al., 2013; Mohseni et al., 2014; Wu et al., 2012).
In a sense, the fact that delocalized Frenkel excitons
can be identified and used to understand spectroscopic
data is a good evidence that the electronic quantum co-
herence, defined here simply as having the quality of co-
herent superposition of site excitation states, plays a sig-
nificant role in the functionality of LHCs. However, this
does not necessarily mean that a specific phase relation-
ship defining a certain exciton state should be preserved,
or change coherently, during a time interval comparable
to that of exciton migration. Rather, the behavior of elec-
tronic quantum coherence can be dynamic or inhomoge-
neous (Ishizaki and Fleming, 2011). These two features
make it extremely difficult to directly detect quantum
coherences that are active in real time through a spec-
troscopic measurement. On the other hand, such features
may in fact play crucial role in positive and cost-effective
utilization of quantum coherence for robust and optimal
exciton dynamics.
In other words, the vast parameter
space of quantum superposition makes it possible for ex-
citons to migrate through different regions while conserv-
ing energy. In such dynamics, quantum coherence offers
(i) redundancy to seek for alternatives when one path is
blocked and (ii) buffering mechanism to help counteract
negative effects of disorder and fluctuations. Thus, gen-
uine understanding of efficient and robust exciton migra-
tion in LHCs seems to require quantitative elucidation
of such redundancy and buffering mechanisms, which in
turn require a holistic approach that considers structure,
energetics, dynamics, disorder, and fluctuations all to-
gether.
The fact that the rate-based ME approach provides a
good description of exciton transfer dynamics in some
LHCs (or some of their parts) does not necessarily mean
that quantum coherence does not play a role as it has
been often assumed. To the contrary, the rate behavior
can be a manifestation of cumulation or averaging of all
possible effects of quantum coherences. In other words,
the extent to which certain signatures of quantum coher-
ences become explicit can be different depending on the
starting point or level of averaging. However, even when
an apparent signature cannot be observed, underneath
the apparent classical-like phenomenology, quantum dy-
namical processes taking advantage of the coherently de-
localized excitons can play important roles.
There are two issues that have significant implications
but were not addressed carefully in this work. One is the
consideration of multiple exciton states and the other
is the dynamics of excitons under natural light condi-
tion. For a complete description of four-wave mixing
spectroscopy and 2DES in particular, it is necessary to
include the double exciton space in order to account for
excited state absorption. Three kinds of double excitons
can be created in general: (i) double excitation of a pig-
ment molecule, (ii) double exciton in the same LHC; (iii)
single excitons in almost two independent LHCs within
the ensemble of spectroscopic measurements. Except for
the trivial case of (iii), calculating energies and couplings
that are relevant to (i) and (ii) are challenging, and can
make the determination of the Hamiltonian much more
complicated. Nonetheless, investigation of the character-
istics of the double excitons and their dynamics have been
shown to be useful even for understanding the behavior
of single excitons (Bruggemann et al., 2001; Trinkunas
et al., 2001).
Another important issue is the understanding of the
nature and the dynamics of excitons under natural inco-
herent light sources from the sun and also how/whether
lessons learned from femtosecond laser spectroscopy can
be used to understand this issue (Brumer and Shapiro,
2012; Chenu et al., 2015, 2014). To this end, formulations
involving excitations by steady incoherent light sources
have been developed (Chenu and Brumer, 2016; Grinev
and Brumer, 2015). Understanding this issue involves
describing photons and their quantum mechanical inter-
actions with nanoscale objects, and can best be studied
through quantum optics experiments of clean and more
well-controlled nanoscale experimental objects.
38
VIII. CONCLUSION AND OUTLOOK
Experimental and theoretical evidences cumulated so
far support that the majority of excitons residing in pig-
ment molecules of LHCs can be well described by the
Frenkel exciton theory. This is in contrast to the excitons
found in most man-made solar energy conversion sys-
tems. For example, excitons created in the first genera-
tion semiconductor-based materials are dominantly Wan-
nier excitons. Even for organic photovoltaic devices con-
sisting of conjugated organic molecules, excitons tend to
have intermediate character and cannot be represented
by simple Frenkel exciton Hamiltonians. Why nature
takes special advantage of Frenkel excitons, which have
been difficult to utilize in man-made systems so far, is
an interesting issue to investigate. A possible explana-
tion based on current experimental and theoretical evi-
dences is that it is related to the properties of proteins
as insulating media and to their superb capability to fine
tune excitation energies and spatial arrangement of pig-
ment molecules. In other words, despite the short coher-
ence length and fragility of Frenkel excitons compared
to Wannier excitons, the versatility of proteins as host-
ing environments allows connecting excitons in energet-
ically and dynamically favorable ways, thereby making
it possible to create efficient and robust exciton-relaying
mechanisms.
The three LHCs reviewed here exemplify different ways
proteins can tune the properties of Frenkel excitons. For
the case of the FMO complex, the main mechanism ap-
pears to be the downhill structure of site excitation ener-
gies through subtle interaction of BChl-protein interac-
tions, whereas weaker electronic couplings and exciton-
bath couplings, also realized through protein's capabil-
ity to control arrangements of BChls, play supporting
roles. For the case of the LH2 complex, circular arrange-
ment of tens of BChls with nearest electronic couplings
on the order of 200 − 300 cm−1 in the B850 unit, main-
tained through stable coordination and hydrogen bonds
to protein residues, seems to be the key feature. Be-
cause the excitation-bath couplings are of moderate mag-
nitudes (on the order of 100 cm−1), the band of dis-
crete Frenkel exciton states spanning about 1, 000 cm−1
are dynamically well connected yet without losing their
identity. Because of alternating in-plane directions of
transition dipoles of α- and β-BChls, the lowest exci-
ton state remains almost dark state, preventing radia-
tive loss through super-radiance. For the case of PBPs,
downhill structure of excitation energies by using differ-
ent pigment molecules and large enough excitation-bath
coupling (Dean et al., 2016; Killoran et al., 2015) real-
ized through covalent bonds, provide efficient pathways
for the migration of Frenkel excitons.
Another important point to consider and, potentially, a
good lesson for developing new generation of photovoltaic
systems, comes from the observation of how LH and CS
domains are put together in natural photosynthetic sys-
tems. In all the cases known so far, the two domains are
always separate, and back exciton transfer pathway from
the CS domain to the LH domain is blocked as much as
possible through ingenious mechanisms. There is also a
"moderating" mechanism of exciton transfer from LH to
RC so as to prevent too fast exciton transfer dynamics
that can damage the CS capability. For purple bacteria,
the moderating function is performed by the size of the
LH1 complex; for green sulfur bacteria, the FMO com-
plex serves such a role by acting as a moderating "valve"
of excitons between chlorosome and reaction centers.
In summary, LHCs are valuable natural systems that
offer great insights into the quantum dynamics of co-
herently delocalized excitons in disordered and insulat-
ing host media. Experimental and theoretical studies
of LHCs involve challenging issues of large scale excited
state quantum dynamics in condensed and complex envi-
ronments, and provide strong motivations and resources
for further advances in these areas of research. Clear
molecular level assessment of the natural design princi-
ples of LHCs should be based on outcomes of these stud-
ies, which in turn can guide the development of genuinely
biomimetic solar light harvesting systems.
ACKNOWLEDGMENTS
SJJ acknowledges the support of the Office of Basic
Energy Sciences, Department of Energy (Grant No. de-
sc0001393) as major sponsor of this research, and the Na-
tional Science Foundation (Grant. No. CHE-1362926) as
the sponsor of research on quantum dynamics methods.
SJJ also thanks Daniel Montemayor and Eva Rivera for
their assistance. BM acknowledges the support of the
European Research Council (ERC) through the Start-
ing Grant, proposal No. 277755 (EnLight). We thank
Greg Scholes, Thomas Renger, Jurgen Kohler, Young
Min Rhee, and David Coker for offering data/images and
for valuable comments. We also thank Tonu Pullerits and
Donatas Zigmantas for discussion and comments.
Appendix A: From molecular Hamiltonian to exciton-bath
Hamiltonian: A review
For the jth pigment molecule, Ψj(cid:105) is defined as the
full and general molecular quantum state. We also ab-
breviate the collective position state of all the nuclear
coordinates of the jth chromophore as Rcj and define
the corresponding position state as follows:
Rcj(cid:105) ≡ Rj,1(cid:105)···Rj,Lj(cid:105) ,
(A1)
39
where the righthand side represents direct product of
states representing Lj (3-dimensional) nuclear coordi-
nates constituting the jth pigment, as has been indicated
in the main text. Then,
(cid:104)Rcj HjΨj(cid:105) =(cid:0)Tj,n(∇cj ) + Vj,nn(Rcj )
(cid:17)(cid:104)RcjΨj(cid:105) ,
+ Hj,e(Rcj )
where
Tj,n(∇cj ) = −
Lj(cid:88)
Lj(cid:88)
l=1
2
2Mj,l
Lj(cid:88)
∇2
j,l ,
(A2)
(A3)
(A4)
1
2
Vj,nn(Rcj ) =
Zj,lZj,l(cid:48)e2
Rj,l − Rj,l(cid:48) ,
Hj,e(Rcj ) = Tj,e + Vj,en(Rcj ) + Vj,ee .
l(cid:48)(cid:54)=l
l=1
(A5)
In the above expression, the definitions of Tj,e and Vj,ee
can be found from Eq. (11). In addition, Vj,en(Rcj ) is a
quantum operator with respect to the electronic degree
of freedom, while depending parametrically on nuclear
coordinates, as follows:
Vj,en(Rcj ) = −
Zj,le2
Rj,l − rj,i .
(A6)
Lj(cid:88)
Nj(cid:88)
l=1
i=1
a. Adiabatic approximation for each chromophore
We denote the αth adiabatic electronic state of the
jth chromophore at Rcj with eigenvalue Ej,α(Rcj ) as
Ej,α(Rcj )(cid:105). Thus,
Hj,e(Rcj )Ej,α(Rcj )(cid:105) = Ej,α(Rcj )Ej,α(Rcj )(cid:105) .
(A7)
Given that electrons are in one of the adiabatic electronic
states, Ej,α(Rcj )(cid:105), for nuclei at Rcj ,
(cid:104)RcjΨj(cid:105) = Φj,α(Rcj )Ej,α(Rcj )(cid:105) ,
(A8)
where Φj,α(Rcj ) is the nuclear wave function for the αth
adiabatic state of the jth pigment. Alternatively, one
can introduce a nuclear state Φj,α(cid:105) such that
Φj,α(Rcj ) = (cid:104)RcjΦj,α(cid:105) .
(A9)
Inserting Eqs. (A8) and (A9) into the right hand side of
Eq. (A2), we obtain
(cid:16)
(cid:17)
Tj,n(∇cj ) + Vj,nn(Rcj ) + Hj,e(Rcj )
= Ej,α(Rcj )(cid:105)(cid:0)Tj,n(∇cj ) + Uj,α(Rcj )(cid:1) Φj,α(Rcj )
×Φj,α(Rcj )Ej,α(Rcj )(cid:105)
Lj(cid:88)
(cid:0)∇j,lΦj,α(Rcj )(cid:1) ·(cid:0)∇j,lEj,α(Rcj )(cid:105)(cid:1)
Lj(cid:88)
2
Mj,l
2
−
l=1
Φj,α(Rcj )∇2
j,lEj,α(Rcj )(cid:105) ,
(A10)
−
2Mj,l
l=1
where Uj,α(Rcj ) is the effective nuclear potential energy
for the αth adiabatic state of the jth pigment. It is the
sum of the nuclear potential energy and the electronic
energy as follows:
resolution for the electronic degrees of freedom can be
expressed as
1jk,e =
Ej,α(cid:105)(cid:104)Ej,α ⊗ Ek,α(cid:48)(cid:105)(cid:104)Ek,α(cid:48)
Uj,α(Rcj ) = Vj,nn(Rcj ) + Ej,α(Rcj ) .
(A11)
=
Ej,α(cid:105)Ek,α(cid:48)(cid:105)(cid:104)Ej,α(cid:104)Ek,α(cid:48) .
(A18)
(cid:88)
(cid:88)
α
(cid:88)
(cid:88)
α(cid:48)
α
α(cid:48)
40
Within the Born-Oppenheimer approximation, the
derivatives of Ej,α(Rcj )(cid:105) with respect to Rcj can be ne-
glected. In addition, assuming that the adiabatic elec-
tronic state is insensitive to the small displacements of
the nuclear coordinate around a reference nuclear coor-
dinate R0
cj
, we can approximate that
Ej,α(Rcj )(cid:105) ≈ Ej,α(R0
cj
)(cid:105) ≡ Ej,α(cid:105) ,
(A12)
where we have assumed that R0
can be defined com-
cj
monly for all adiabatic states, namely, independent of α,
for the jth pigment.
Taking the inner product of (cid:104)Ej,α with Eq. (A2), em-
ploying Eqs. (A8) and (A10), and applying the Born-
Oppenheimer approximation, we obtain
(cid:104)Ej,α(cid:104)Rcj HjΨj(cid:105)
= (Tj,n(∇cj ) + Uj,α(Rcj ))Φj,α(Rcj )
= (cid:104)Rcj( Tj,n + Uj,α( Rcj ))Φj,α(cid:105) .
(A13)
The above relation holds for any value of Rcj (within
the approximation of Eq. (A12)), and thus leads to the
following identity:
(cid:104)Ej,α HjΨj(cid:105) = ( Tj,n + Uj,α( Rcj ))Φj,α(cid:105) . (A14)
Applying the same procedure as described above for all
other adiabatic electronic states and considering linear
combination of them, we can express the general molec-
ular quantum state of the jth chromophore as follows:
Cj,αΦj,α(cid:105)Ej,α(cid:105) ,
(A15)
(cid:88)
malization condition (cid:80)
Ψj(cid:105) =
α
where Cj,α's are complex coefficients satisfying the nor-
α Cj,α2 = 1. Applying Hj to
Eq. (A15) and invoking the Born-Oppenheimer approx-
imation for each adiabatic electronic state, we obtain
HjΨj(cid:105) =
Cj,α( Tj,n+Uj,α( Rcj ))Φj,α(cid:105)Ej,α(cid:105) . (A16)
(cid:88)
α
This expression is equivalent to the following representa-
tion of the Hamiltonian of the jth pigment molecule:
(cid:88)
( Tj,n + Uj,α( Rcj ))Ej,α(cid:105)(cid:104)Ej,α .
Hj =
(A17)
α
The set of Ej,α(cid:105)'s forms a complete basis and constitutes
the site basis of the Frenkel exciton states.
In the direct product space of the electronic states con-
stituting the jth and kth pigment molecules, the identity
function:
ρα(cid:48)(cid:48)α(cid:48)
j
(rj,1) = Nj
Each term constituting Hjk of Eq. (12) can be projected
into this basis of electronic states. The first term rep-
resenting nuclear-nuclear repulsion, Vjk,nn, remains the
same and diagonal in the electronic basis. The second
term representing the attraction between electrons in the
jth pigment and nuclei in the kth pigment can be ex-
pressed as
Vjk,ne =
V α(cid:48)α(cid:48)(cid:48)
jk,ne ( Rcj )
(cid:88)
(cid:88)
α
α(cid:48),α(cid:48)(cid:48)
×Ej,α(cid:105)Ek,α(cid:48)(cid:105)(cid:104)Ej,α(cid:104)Ek,α(cid:48)(cid:48) ,
(A19)
where
V α(cid:48)α(cid:48)(cid:48)
jk,ne ( Rcj )
= −
Lj(cid:88)
Nk(cid:88)
(cid:104)Ek,α(cid:48)
l=1
i=1
Zj,le2
Rj,l − rk,iEk,α(cid:48)(cid:48)(cid:105) .
(A20)
Similarly, the third term representing the attraction be-
tween electrons of the jth pigment and the nuclei of the
kth pigment can be expressed as
(cid:88)
(cid:88)
Vjk,en =
V αα(cid:48)
jk,en( Rck )
α(cid:48)(cid:48)
α,α(cid:48)
×Ej,α(cid:105)Ek,α(cid:48)(cid:48)(cid:105)(cid:104)Ej,α(cid:48)(cid:104)Ek,α(cid:48)(cid:48) ,
where
(A21)
(A22)
V αα(cid:48)
jk,en( Rck )
= − Lk(cid:88)
l=1
Nj(cid:88)
(cid:104)Ej,α Zk,le2
(cid:88)
(cid:88)
i=1
Vjk,ee =
V αα(cid:48),α(cid:48)(cid:48)α(cid:48)(cid:48)(cid:48)
jk,ee
Rk,l − rj,iEj,α(cid:48)(cid:105) .
Finally, the fourth term representing repulsion between
electrons of the jth and kth pigment can be expressed as
α(cid:48)(cid:48),α(cid:48)(cid:48)(cid:48)
α,α(cid:48)
Ej,α(cid:105)Ek,α(cid:48)(cid:48)(cid:105)(cid:104)Ej,α(cid:48)(cid:104)Ek,α(cid:48)(cid:48)(cid:48) ,
(A23)
where
V αα(cid:48),α(cid:48)(cid:48)α(cid:48)(cid:48)(cid:48)
jk,ee
Nj(cid:88)
Nk(cid:88)
=
i=1
i(cid:48)=1
(cid:104)Ej,α(cid:104)Ek,α(cid:48)(cid:48)
e2
rj,i − rk,i(cid:48)Ej,α(cid:48)(cid:105)Ek,α(cid:48)(cid:48)(cid:48)(cid:105) .
(A24)
We here define the following electron density matrix
(cid:90) Nj(cid:89)
(cid:104)Ej,α(cid:48)rcj(cid:105)(cid:104)rcjEj,α(cid:48)(cid:48)(cid:105) ,
drj,i
i=2
(A25)
where rcj refers to the collection of all the electron co-
ordinates of the jth pigment. For α(cid:48) = α(cid:48)(cid:48), this cor-
responds to the single electron density at rj,1. On the
other hand, for α(cid:48) (cid:54)= α(cid:48)(cid:48), it becomes the transition den-
sity. All the interaction terms between different pigment
molecules shown above can then be expressed in terms of
the above electron density matrix. First, Eq. (A20) can
be shown to be
V α(cid:48)α(cid:48)(cid:48)
jk,ne ( Rcj )
= −
Nk(cid:88)
(cid:104)Ek,α(cid:48)rck(cid:105)(cid:104)rck
Lj(cid:88)
Rj,l − rk,iEk,α(cid:48)(cid:48)(cid:105)
Zj,le2
drck
l=1
i=1
dr ρα(cid:48)(cid:48)α(cid:48)
k
(r)
Zj,le2
Rj,l − r .
Lk(cid:88)
Nj(cid:88)
(cid:104)Ej,αrcj(cid:105)(cid:104)rcj Zk,le2
Rk,l − rj,iEj,α(cid:48)(cid:105)
(cid:90)
Lj(cid:88)
l=1
(cid:90)
= −
Similarly,
V αα(cid:48)
jk,en( Rck )
= −
(cid:90)
= − Lk(cid:88)
(cid:90)
drcj
l=1
l=1
i=1
dr ρα(cid:48)α
j
(r)
(cid:90)
Zk,le2
Rk,l − r .
(A27)
On the other hand, the off-diagonal elements are given
by
(A26)
(cid:104)sj Hcsj(cid:105) = (cid:104)g Hcg(cid:105) + Uj,1( Rcj ) − Uj,0( Rcj )
41
where the product is over all k = 1,··· , Nc except for j.
The molecular Hamiltonian for the aggregate of pig-
ments, Hc, can be expressed in the basis spanned by g(cid:105)
and sj(cid:105) as defined above by employing the expressions
obtained in the previous subsection. First, the diagonal
elements of Hc in the site excitation basis are given by
(cid:104)g Hcg(cid:105) =
(cid:17)
(cid:16) Tj,n + Uj,0( Rcj )
Nc(cid:88)
(cid:16)
Nc(cid:88)
(cid:88)
j=1
Vjk,nn( Rcj , Rck ) + V 00
jk,ne( Rcj )
+
1
2
j=1
k(cid:54)=j
(cid:17)
+V 00
jk,en( Rck ) + V 00,00
jk,ee
,
(A31)
(cid:16)
Nc(cid:88)
k(cid:54)=j
+
1
2
(cid:104)sj Hcg(cid:105) =
(cid:104)g Hcsj(cid:105) =
kj,ne( Rcj )
jk,en( Rcj )
kj,ee + V 11,00
+V 11
+V 00,11
kj,ne( Rcj ) − V 00
V 11
jk,en( Rcj ) − V 00
kj,ee − V 00,00
(cid:16)
Nc(cid:88)
(cid:16)
Nc(cid:88)
k(cid:54)=j
1
2
1
2
k(cid:54)=j
kj,ne( Rck ) + V 10
V 10
jk,ee
jk,ee − V 00,00
(cid:17)
(cid:17)
jk,en( Rck )
kj,ne( Rck ) + V 01
V 01
jk,en( Rck )
(cid:17)
. (A32)
= J c,0
jg
(A33)
= J c,0
gj
(A34)
jk,ee = J c,0
jk , for j (cid:54)= k .
(cid:104)sj Hcsk(cid:105) = V 10,01
In Eqs. (A33) and (A34), the fact that V 00,10
jk,ee =
V 00,01
kj,ee = V 01,00
jk,ee = 0 has been used, which can be proved
from the fact that the integration of the transition density
over the entire electronic coordinates becomes zero.
kj,ee = V 10,00
(A35)
Let us define
Nc(cid:88)
Nc(cid:88)
j=1
j=1
k(cid:54)=j
Uj,0( R0
cj
)
(cid:88)
(cid:16)
Ec,0
g =
+
1
2
Vjk,nn( R0
cj
, R0
ck
) + V 00
jk,ne( R0
cj
)
+V 00
) + V 00,00
jk,ee
,
(A36)
(cid:17)
jk,en( R0
ck
) − Uj,0( R0
) − V 00
)
cj
Ec,0
j = Ec
g + Uj,1( R0
cj
Nc(cid:88)
(cid:16)
k(cid:54)=j
+V 11
+
1
2
kj,ne( R0
V 11
ck
kj,ne( R0
ck
)
jk,en( R0
ck
kj,ee − V 00,00
) − V 00
jk,en( R0
ck
kj,ee + V 11,00
+V 00,11
)
jk,ee − V 00,00
jk,ee
(cid:17)
.(A37)
Finally, it is easy to show that
V αα(cid:48),α(cid:48)(cid:48)α(cid:48)(cid:48)(cid:48)
jk,ee
=
drdr(cid:48)ρα(cid:48)α
j
(r)ρα(cid:48)(cid:48)(cid:48)α(cid:48)(cid:48)
k
(r(cid:48))
e2
r − r(cid:48) . (A28)
This corresponds to the Coulomb interaction term be-
tween α → α(cid:48) electronic transition of the jth pigment
and α(cid:48)(cid:48) → α(cid:48)(cid:48)(cid:48) electronic transition of the kth pigment.
b. Hamiltonian in the site excitation basis
The site basis for the electronic states of the aggre-
gates of pigments can be constructed by taking direct
product of all the adiabatic electronic states. For the
sake of simplicity, we here assume that the index α can
be represented by nonnegative integers and there is no
degeneracy in the ground and the first excited adiabatic
electronic states of each pigment. Thus, α = 0 represents
the ground electronic state and α = 1 the first excited
state of each pigment. First, the ground electronic state
of the aggregates can be expressed as
g(cid:105) =
Ej,0(cid:105) .
(A29)
Nc(cid:89)
j=1
The state where only the jth pigment is excited (site
excitation state) is defined as
(cid:89)
k(cid:54)=j
Ej,1(cid:105) ,
sj(cid:105) =
Ek,0(cid:105)
(A30)
Collecting all the terms involving nuclear degrees of free-
dom, we can also define
Nc(cid:88)
Nc(cid:88)
j=1
(cid:16) Tj,n + Uj,0( Rcj ) − Uj,0( R0
(cid:16)
(cid:88)
)
cj
(cid:17)
j=1
k(cid:54)=j
Vjk,nn( Rcj , Rck ) − Vjk,nn( R0
(cid:17)
jk,ne( Rcj ) − V 00
jk,en( Rck ) − V 00
jk,ne( R0
cj
jk,en( R0
ck
c,j = Uj,1( Rcj ) − Uj,0( Rcj ) − Uj,1( R0
) + Uj,0( R0
B0
cj
, R0
ck
)
+V 00
+V 00
cj
cj
)
)
,
)
(A38)
H 0
c,b =
+
1
2
(cid:16)
Nc(cid:88)
k(cid:54)=j
+
1
2
kj,ne( Rck ) − V 11
V 11
kj,ne( R0
ck
)
−V 00
+V 11
−V 00
kj,ne( Rck ) + V 00
jk,en( Rck ) − V 11
jk,en( Rck ) + V 00
kj,ne( R0
ck
jk,en( R0
ck
jk,en( R0
ck
)
)
)
(cid:17)
.
(A39)
Collecting all these terms, we obtain the exciton-bath
Hamiltonian defined by Eqs. (13). It is important to note
that the ground electronic state and the single exciton
state in this section has been defined as direct product
of those of independent pigment molecules. In practice,
these can be improved further by defining them as those
accounting for the implicit effect of the other pigment
molecules in the ground electronic states.
Appendix B: Modeling of Environmental effects
In the context of the modeling of electronic processes in
embedded systems, the most used multiscale approach is
that combining the QM description of the chromophores
with a classical description of the environment. The suc-
cess of such QM/classical formulation is mostly related
to the accuracy with which the interactions between the
QM and the classical parts are treated. Generally speak-
ing, we can identify two different classes of QM/classical
formulations, namely that describing the environment
as a continuum, in terms of its macroscopic properties
(QM/continuum), and that keeping an atomistic descrip-
tion through MM force fields (QM/MM). Both methods
allow to account for the presence of the environment in
the description of a molecular system at an affordable
computational cost, increasing both the possibilities of
molecular modeling and the manifold of treatable sys-
tems. In both versions of QM/classical models, an im-
portant common aspect is that the QM part can be mod-
ified in its electronic and nuclear degrees of freedom by
the presence of the classical part. From a QM point of
view, the Hamiltonian of the whole system can be written
as:
H = HQM + Henv + Hint
(B1)
42
where HQM is the Hamiltonian of the gas-phase (iso-
lated) QM subsystem, Henv is the Hamiltonian of the
rest of the system, which is purely classical, and finally
Hint represents the interaction between the QM and the
classical parts. In the QM/classical models the degrees
of freedom of the sole environment are not of great rele-
vance. In continuum solvation models, the solvent atom-
istic nature disappears together with the Henv term. In
the case of QM/MM approach, Henv is maintained, but
such term only adds a constant term to the total energy.
In continuum models, the QM subsystem is placed in
a suitably shaped molecular cavity C immersed in the di-
electric medium representing the environment. The po-
larization response of the medium to the QM charge dis-
tribution is obtained by solving the Poisson equation of
classical electrostatics (Tomasi et al., 2005). If we assume
that the charge distribution (ρ) of the QM subsystem is
entirely contained inside the cavity and that the dielectric
outside the cavity is characterized by a scalar dielectric
constant, we obtain
(cid:40)−∇2V (r) = 4πρ(r) within the cavity
−∇2V (r) = 0
outside the cavity
(B2)
A possible strategy to solve this equation is to write the
electrostatic potential V as a sum of the solute potential
plus the contribution due to the reaction of the environ-
ment (i.e.
the polarization of the dielectric). Among
the possible approaches to define the reaction poten-
tial, a very effective one is the apparent surface charges
(ASC) approach, where an apparent surface density σ(s)
spread on the cavity surface Γ is used to represent the
reaction (polarization) of the dielectric (Tomasi et al.,
2005). From a computational point of view, the solution
is achieved by partitioning the cavity surface into a set of
finite elements and substituting σ(s) with a set of point
charges (qt), each corresponding to a surface element.
There have been different definitions of qt proposed,
leading to different formulations of the so-called Po-
larizable Continuum Models (PCMs)(Mennucci, 2012),
or to Conductor-like Screening Model (COSMO)(Klamt,
2011). Within this framework, the operator Hint in Eq.
(B1) represents the electrostatic interaction between the
QM charge density and the apparent charges on the sur-
face of the cavity:
Hint = HQM/PCM =
qt VQM(rt) ,
(B3)
t
where qt is the ASC at the position rt in the Cartesian
coordinate system, and VQM (rt) is the electrostatic po-
tential operator due to the QM charge density calculated
at rt. The summation runs over all the Nt surface ele-
ments. Since HQM/PCM depends on the QM charge den-
sity which is modified by the environment through Hint,
a non-linear problem is obtained and its solution leads to
a mutually polarized QM/continuum system.
(cid:88)
Continuum models are often the ideal strategy to ac-
count for the bulk effects of the environment. On the
other hand, the use of a structureless medium, does not
allow to include the effects of the heterogeneity, as well
as specific interactions between the two parts.
In par-
ticular, when the heterogeneity of the environment acts
at a small scale as in a protein matrix, where each po-
sition/orientation of the QM subsystem feels a different
local environment or in the presence of specific interac-
tions such as hydrogen bonds, an atomistic description
is preferred. In these cases, QM/MM models represent
a very effective strategy. However, they also suffer from
a drawback: the QM-environment interaction depends
critically on the configuration of the environment, par-
ticularly those at short range. Therefore, to correctly
account for the dynamical nature of the interactions, sev-
eral configurations of the whole system need to be taken
into consideration. This is not needed when a continuum
approach is employed since it implicitly gives a configu-
rationally sampled effect due to the use of macroscopic
properties. Indeed, the sampling issue may introduce an
additional difficulty related to the fact that it often in-
volves classical MD simulation, which can incur errors
resulting from the inaccuracy of the force-field used and
the intrinsic approximation of classical MM simulations.
In the most common formulation, the MM part of
QM/MM systems is described through a set of fixed
point charges, generally placed at the atoms of the envi-
ronment. The QM/classical interactions are therefore of
electrostatic nature. The MM charges are chosen to best
represent the molecular electrostatic properties. Eventu-
ally, higher multipole moments can be used to improve
the electrostatic description.
The Hint term in Eq. (B1) is given by the interaction
between the electronic density of the QM subsystem ρs(r)
and the MM charge distribution ( HQM/MM). Namely,
Hint = HQM/MM =
qm VQM(rm) ,
(B4)
m
where qm's are the fixed partial charges placed at position
rm's in the Cartesian coordinate system and VQM (rm) is
the electrostatic potential operator due to the QM charge
density calculated at rm. The summation runs over all
the number of charges. This scheme is also known as
"electrostatic embedding" in the sense that the electronic
structure of the QM part is modified by the presence of
the charge distribution representing the environment and
consequentially is polarized by it. However, the use of
fixed point charges to describe the MM system implies
that, while the QM density is polarized by the MM one,
the opposite is not true. This is a serious limitation, as
the explicit response arising from the environment po-
larization can be crucial, particularly when charged or
very polar systems are studied, or when electronic ex-
citation processes are considered.
In order to improve
the QM/MM description, a possible strategy is to use
(cid:88)
43
a "flexible" MM model which can be polarized back by
the QM charge distribution. Three main groups of meth-
ods include these mutual polarization effects (Mennucci,
2013; Senn and Thiel, 2009). In the induced dipole (ID)
approach, atomic polarizabilities are assigned to atoms
that lead to induced dipoles in the presence of an elec-
tric field. The source of the electric field are the QM
charge distribution, the MM point charges and the in-
duced dipoles themselves. A self-consistent procedure
is required to solve the QM/MM problem because the
dipoles interact with each other and act back on the QM
electron density. The parameters to be defined in addi-
tion to the MM charges are the atomic polarizabilities.
If Drude oscillators (DO) are used, instead, atoms are
represented by a pair of point charges separated by a
variable distance (namely a spring). When interacting
with an electric field, a dipole is generated. The pa-
rameters are the magnitude of the mobile charge and the
force constant of the spring. Finally, when the fluctuating
charges (FQ) model is used, charges placed on the atoms
are allowed to fluctuate so as to represent the charge flow
within the molecule. Two sets of atomic parameters are
needed (hardness and electronegativities) as the model
is based on the electronegativity equalization principle,
which states that, at equilibrium, the instantaneous elec-
tronegativity of each atom has the same value.
In the ID approach, a set of atomic polarizabilities is
assigned to MM atoms. The polarizability αi is a 3 × 3
tensor. However, in most formulations of the model, only
isotropic component αiso is used. Within this framework,
the induced dipole moment µ at the MM site i can be
written as
Eext
i −
Np(cid:88)
j(cid:54)=i
µi = αiso
i
Tijµj
(B5)
i
where Eext
is the electric field due to the QM subsystem
and the set of MM atomic charges. The Tij is known as
the dipole-dipole interaction tensor and is defined as:
r2
Tij =
fe
r3
ij
I − 3ft
r5
ij
rxry rxrz
x
r2
ryrz
ryrx
y
r2
rzrx rzry
z
(B6)
where I is the 3 x 3 unit tensor and rx, ry and rz are
Cartesian components of the vector connecting the two
atoms i and j. The fe and ft factors are distance-
dependent screening functions that depend on the spe-
cific dipole interaction models.
In recent years, atomistic (MM) and continuum ap-
proaches have been coupled to give fully polarizable
QM/MM/continuum approaches (Boulanger and Thiel,
2012; Caprasecca et al., 2012; Lipparini and Barone,
2011; Steindal et al., 2011). In these methods, on top of
the polarizable discrete model, one adds a further exter-
nal polarizable continuum layer, coupled to the former.
In this way the continuum description of the environment
is combined together with the polarizable QM/MM, thus
obtaining a three level model which allows to exploit the
advantages of each method. The polarizable MM model
is used to describe short-range directional interactions,
whereas the continuum one accounts for long-range (or
bulk) effects.
When hybrid QM/classical models are used to describe
ultrafast processes such as electronic excitations in em-
bedded chromophores, a new aspect regarding the char-
acteristic response time of the environmental degrees of
freedom has to be taken into account. When a ver-
tical excitation occurs,
in fact, the electronic (or dy-
namic) component of the response, which is assumed to
be sufficiently fast, immediately follows any change in
the chromophore charge distribution. On the contrary,
the slower (or inertial) response that arises from nuclear
and molecular motions remains frozen in its initial state.
The possible delay of the slower component results in
a nonequilibrium regime, which eventually relaxes into
a new equilibrium where all the degrees of freedom of
the environment reach the equilibrium with the excited
state chromophore. Especially for highly polar environ-
ments, equilibrium and nonequilibrium regimes represent
very different configurations and their energy difference
is generally known as "reorganization energy."
The nonequilibrium regime can be properly taken into
account with both the QM/continuum and QM/MM
models.
In the first case, if we adopt the ASC frame-
work, the fast (or dynamic) response will be represented
in terms of apparent charges obtained from the the opti-
cal component (∞) of the dielectric permittivity instead
of the static one. By contrast, the slow response is ob-
tained as the difference of the equilibrium charges and
the dynamic ones. In the QM/MM framework, instead,
the slow component is automatically taken into account
if the MM charge distribution does not change during
the excitation process (charges are fixed in position and
in value). For what concerns the fast component, this
can be properly taken into account only if a polarizable
embedding is considered.
Appendix C: Lineshape and response functions
The model considered here is more general than what
is considered in the main text because we assume here
that the EBH can have additional time dependence due
to fluctuating energies and couplings. Thus, the total
Hamiltonian including interaction with the radiation has
the following form:
44
Nc(cid:88)
Nc(cid:88)
where Hint(t) is the matter-radiation interaction Hamil-
tonian and
Hex(t) = He(t) +
Bjksj(cid:105)(cid:104)sk + Hb ,
(C2)
j=1
k=1
with
He(t) =
Nc(cid:88)
j=1
Ej(t)sj(cid:105)(cid:104)sj +
Nc(cid:88)
(cid:88)
j=1
k(cid:54)=j
Jjk(t)sj(cid:105)(cid:104)sk . (C3)
The time dependences in Eg(t), Ej(t), and Jjk(t) repre-
sent the influence of all other sources that are not rep-
resented by the bath Hamiltonian. The details of these
time dependent terms depend on specific experimental
conditions under which a specific spectroscopic measure-
ment is made.
The eigenstates of He(t) are denoted as ϕj(t)(cid:105)'s, and
we introduce Mkj(t) = (cid:104)skϕj(t)(cid:105). Thus,
kj(t)ϕj(t)(cid:105) .
(cid:88)
sk(cid:105) =
M∗
(C4)
j
The transition dipole vector for the excitation from g(cid:105)
to sk(cid:105), µk, is assumed to be time dependent in general.
Then, the polarization operator for the transitions to the
single exciton space at time t is given by
µk(t) (sk(cid:105)(cid:104)g + g(cid:105)(cid:104)sk)
(Dj(t)ϕj(t)(cid:105)(cid:104)g + g(cid:105)(cid:104)ϕj(t)Dj(t)) ,(C5)
(cid:88)
(cid:88)
where Dj(t) =(cid:80)
P(t) =
=
k
j
kj(t).
The time evolution operator for the total Hamiltonian
Eq. (C1) is denoted as
k µk(t)M∗
(cid:26)
(cid:90) t
t0
(cid:27)
UT (t, t0) = exp(+)
− i
dt(cid:48) HT (t(cid:48))
,
(C6)
where (+) represents chronological time ordering.
Thus, given that the total density operator at time t0
is prepared to be ρ(t0), at time t, it evolves into
ρ(t) = UT (t, t0)ρ(t0) U†
T (t, t0) .
(C7)
For the discussion that follows, it is convenient to in-
troduce the total Hamiltonian (without matter-radiation
interaction) in the ground electronic state and the man-
ifold of single exciton states as follows:
Hg(t) = Eg(t)g(cid:105)(cid:104)g + Hb ,
H(t) = Eg(t)g(cid:105)(cid:104)g + Hex(t) .
(C9)
(C8)
We also denote the collection of all the time dependent
parameters as
HT (t) = Eg(t)g(cid:105)(cid:104)g + Hex(t) + Hint(t) ,
(C1)
Γ(t) ≡ (Eg(t), Ej(t)'s, Jjk(t)'s, µk(t)'s)
(C10)
Ie =
sj(cid:105)(cid:104)sj ,
2. Emission
(C14)
1. Absorption
For t ≤ t0, the system is in the ground electronic state
and the bath is in thermal equilibrium. Therefore,
ρ(t0) = g(cid:105)(cid:104)gρb ,
(C11)
where ρb = e−β Hb /T rb{e−β Hb} with β = 1/kBT . For
a monochromatic radiation with frequency ω and polar-
ization η, the matter-radiation interaction Hamiltonian
(within the semiclassical approximation for the radiation
and the rotating wave approximation) can be expressed
as
Hint(t) = Ae−iωtD(t)(cid:105)(cid:104)g + A∗eiωtg(cid:105)(cid:104)D(t) ,
(C12)
where A is the amplitude of the radiation including the
wave vector, and
D(t)(cid:105) =
η · Dj(t)ϕj(t)(cid:105) .
(C13)
Expanding Eq. (C7) up to the second order of Hint(t)
and introducing the following identity operator in the
single exciton space,
(cid:88)
j
(cid:88)
j
the probability to find the population of exciton at time
t, when approximated up to the second order of matter-
radiation interaction, becomes
(cid:90) t
dt(cid:48)(cid:90) t
dt(cid:48)(cid:90) t
t0
t0
1
2
(cid:90) t
(cid:110) Ie U(t, t(cid:48)) Hint(t(cid:48))
(cid:111)
(cid:110) Ug(t0, t(cid:48)(cid:48))g(cid:105)(cid:104)D(t(cid:48)(cid:48))
Pe(t) =
dt(cid:48)(cid:48) T r
× U(t(cid:48), t0)ρ(t0) U†(t(cid:48)(cid:48), t0) Hint(t(cid:48)(cid:48)) U†(t, t(cid:48)(cid:48))
=
A2
2
× Uex(t(cid:48)(cid:48), t(cid:48))D(t(cid:48))(cid:105)(cid:104)g Ug(t(cid:48), t0)ρ(t0)
dt(cid:48)(cid:48)eiω(t(cid:48)(cid:48)−t(cid:48))T r
t0
t0
(cid:111)
(cid:27)
dt(cid:48) H(t(cid:48))
− i
(C16)
dτ Eg(τ )g(cid:105)(cid:104)ge− i Hb(t−t0) , (C17)
,
(cid:27)
dt(cid:48) Hex(t(cid:48))
.
(C18)
where
(cid:90) t
(cid:90) t
t0
t0
− i
(cid:26)
U(t, t0) = exp(+)
− i(cid:82) t
Ug(t, t0) = e
Uex(t, t0) = exp(+)
(cid:90) t
(cid:26)
t0
A2
2 Re
Taking time derivative of Eq. (C15), one obtains
Pe(t) = 2
d
dt
× g(cid:105)(cid:104)D(t) Uex(t, t(cid:48))D(t(cid:48))(cid:105)(cid:104)g Ug(t(cid:48), t0)ρ(t0)
dt(cid:48)eiω(t−t(cid:48))T r
t0
.(C19)
(cid:110) Ug(t0, t)
(cid:111)
45
Let us introduce the following "ideal" lineshape by di-
viding the above time derivative with appropriate nor-
malization factor:
Iid[ω, Γ(t)] =
2
Pe(t)
2πA2
d
dt
dt(cid:48) eiω(t−t(cid:48))e
(cid:90) t
i(cid:82) t
(cid:110) Uex(t, t(cid:48))D(t(cid:48))(cid:105)(cid:104)D(t)ρbe
t0
Re
1
=
π
×T r
t(cid:48) dτ(cid:48)Eg(τ(cid:48))
i Hb(t−t(cid:48))(cid:111)
. (C20)
In general, the absorption line shape can be defined as
I(ω) = lim
t→ts
(cid:104)Iid[ω; Γ(t)](cid:105)Γ(t) .
(C21)
For the simple case where there is no time dependent
fluctuation of Eg(t) and parameters constituting Hex(t),
Eq. (C20) reduces to
(cid:110)
Iid(ω) =
×T r
1
π
Re
0
dτ eiωτ e
i τ Eg
(cid:111)
e− i τ HexD(cid:105)(cid:104)Dρbe
i τ Hb
(C22)
Taking the limit of t − t0 → ∞ and making average over
the disorder, orientation, and polarization leads to Eq.
(49).
.
(cid:90) t−t0
Assume that the matter is in single exciton space and
is in equilibrium at time t0. Then, the initial density
operator can be expressed as
ρ(t0) =
e−β Hex(t0)
T r{e−β Hex(t0)} .
(C23)
For the above initial density operator, following a pro-
cedure similar to obtaining Eq. (C15), the population
of the ground state at time t, when approximated up to
the second order of matter-radiation interaction, can be
shown to be
(cid:110)g(cid:105)(cid:104)g U(t, t(cid:48)) Hint(t(cid:48))
(cid:111)
(cid:110) Uex(t0, t(cid:48)(cid:48))D(t(cid:48)(cid:48))(cid:105)(cid:104)g
× U(t(cid:48), t0)ρ(t0) U†(t(cid:48)(cid:48), t0) Hint(t(cid:48)(cid:48)) U†(t, t(cid:48)(cid:48))
=
A2
2
× Ug(t(cid:48)(cid:48), t(cid:48))g(cid:105)(cid:104)D(t(cid:48)) Uex(t(cid:48), t0)ρ(t0)
dt(cid:48)(cid:48)eiω(t(cid:48)−t(cid:48)(cid:48))T r
t0
t0
(cid:111)
.
(C24)
(cid:90) t
dt(cid:48)(cid:90) t
dt(cid:48)(cid:90) t
t0
t0
1
2
(cid:90) t
Then,
2
Eid[ω, Γ(t)] =
d
dt
Pg(t)
2πA2
dt(cid:48) e−iω(t−t(cid:48))e− i(cid:82) t
(cid:90) t
(cid:110) Uex(t0, t)e− i (t−t(cid:48)) HbD(t)(cid:105)(cid:104)D(t(cid:48)) Uex(t(cid:48), t0)ρ(t0)
(cid:111)
t(cid:48) dτ(cid:48)Eg(τ(cid:48))
t0
Re
1
=
π
×T r
(C25)
.
(C15)
Pg(t) =
dt(cid:48)(cid:48) T r
In general, the emission line shape can be defined as
where
E(ω) = lim
t→ts
(cid:104)Eid[ω; Γ(t)](cid:105)Γ(t) .
(C26)
For the simple case where there is no time dependent
fluctuation of Eg(t) and parameters constituting Hex(t),
Eq. (C25) reduces to
(cid:110)
Eid[ω, Γ(t)] =
×T r
1
π
Re
0
dτ e−iωτ e− i τ Eg
(cid:111)
e− i τ HbD(cid:105)(cid:104)Dρexe
i τ Hex
,
(C27)
(cid:90) t−t0
where ρex = e−β Hex/T r{e−β Hex}. Taking the limit of
t − t0 → ∞ and making average over the disorder, orien-
tation, and polarization of the above expression leads to
Eq. (50).
46
(C33)
dt
dt(cid:48)(cid:48) U(tm, t)
ρI (tm) = − i
3
× Hint(t) U(t, t(cid:48)) Hint(t(cid:48)) U(t(cid:48), t0)ρ(t0)
× U†(t(cid:48)(cid:48), t0) Hint(t(cid:48)(cid:48)) U†(tm, t(cid:48)(cid:48)) ,
0
0
(cid:90) tm
(cid:90) tm
(cid:90) t
(cid:90) t
dt(cid:48)(cid:90) tm
dt(cid:48)(cid:90) t(cid:48)
0
0
0
0
dt
dt(cid:48)(cid:48) U(tm, t0)ρ(t0)
ρII (tm) = − i
3
× U†(t(cid:48)(cid:48), t0) Hint(t(cid:48)(cid:48)) U†(t(cid:48), t(cid:48)(cid:48)) Hint(t(cid:48))
× U†(t, t(cid:48)) Hint(t) U†(tm, t) .
(C34)
In Eq. (C33), the integration over t(cid:48)(cid:48) can be split into
three regions, 0 < t(cid:48)(cid:48) < t(cid:48), t(cid:48) < t(cid:48)(cid:48) < t, and t < t(cid:48)(cid:48) < tm.
Relabeling the dummy time integration variables in each
region such that t ≥ t(cid:48) ≥ t(cid:48)(cid:48), the three terms can be
rewritten so as to have the same time integration bound-
aries as ρII (tm). The resulting third order components
can therefore be expressed as
(cid:90) tm
(cid:90) t
dt(cid:48)(cid:90) t(cid:48)
dt(cid:48)(cid:48) 4(cid:88)
dt
0
0
0
j=1
Tj(tm, t, t(cid:48), t(cid:48)(cid:48))
(C35)
T1(tm, t, t(cid:48), t(cid:48)(cid:48)) ≡ U(tm, t(cid:48)) Hint(t(cid:48)) U(t(cid:48), t(cid:48)(cid:48)) Hint(t(cid:48)(cid:48))
× U(t(cid:48)(cid:48), t0)ρ(t0)U†(t, t0) Hint(t)U†(tm, t) ,
T2(tm, t, t(cid:48), t(cid:48)(cid:48)) ≡ U(tm, t) Hint(t) U(t, t(cid:48)(cid:48)) Hint(t(cid:48)(cid:48))
× U(t(cid:48)(cid:48), t0)ρ(t0) U†(t(cid:48), t0) Hint(t(cid:48)) U†(tm, t(cid:48)) ,
T3(tm, t, t(cid:48), t(cid:48)(cid:48)) ≡ U(tm, t) Hint(t) U(t, t(cid:48)) Hint(t(cid:48))
× U(t(cid:48), t0)ρ(t0) U†(t(cid:48)(cid:48), t0) Hint(t(cid:48)(cid:48)) U(tm, t(cid:48)(cid:48)) ,
T4(tm, t, t(cid:48), t(cid:48)(cid:48)) ≡ U(tm, t0)ρ(t0) U†(t(cid:48)(cid:48), t0) Hint(t(cid:48)(cid:48))
× U†(t(cid:48), t(cid:48)(cid:48)) Hint(t(cid:48)) U†(t, t(cid:48)) Hint(t) U†(tm, t) .
(C36)
(C37)
(C38)
(C39)
The corresponding third order contribution to the po-
larization can be calculated by taking the trace of the
scalar product between P(tm), Eq. (C5), and ρ(3)(tm),
Eq. (C35). The resulting expression for the third order
polarization at time tm can be shown to be
4(cid:88)
(cid:110) P(tm) Tj(tm, t, t(cid:48), t(cid:48)(cid:48))
(cid:111)
2
3 Im
j=1
(cid:90) tm
dt
0
,
(C40)
¯P(3)(tm) ≡ T r{ P(tm)ρ(3)(tm)} =
(cid:90) t
dt(cid:48)(cid:90) t(cid:48)
×
dt(cid:48)(cid:48)T r
0
0
where "Im" implies imaginary part of the complex func-
tion.
Let us assume that we are interested in the polarization
along the direction of ηm at time tm, and also define
Dm(tm)(cid:105) =
ηm · Dj(t)ϕj(t)(cid:105) ,
j
Pmg(tm) = Dm(tm)(cid:105)(cid:104)g .
(C41)
(C42)
(cid:88)
3. Four wave mixing spectroscopy
ρ(3)(tm) = − i
3
For three incoming pulses, the matter-radiation inter-
action Hamiltonian is given by
+H.c. ,
where
3(cid:88)
(cid:88)
Hint(t) =
Eν(t − tν) · Dj(t)ϕj(t)(cid:105)(cid:104)geikν·r−iων t
j
ν=1
+H.c.
Eν(t − tν)Dν(t)(cid:105)(cid:104)geikν·r−iων t + H.c. .
(C28)
3(cid:88)
ν=1
=
In the above expression, Eν(t − tν) = Eν(t − tν)ην, with
Eν(t−tν) and ην respectively defining the amplitude and
unit polarization vector of the νth pulse. It is assumed
that t3 ≥ t2 ≥ t1. In the last line of Eq. (C28), Dν(t)(cid:105)
is the sum of all the exciton states weighted by the com-
ponents of the transition dipoles along the direction ην
and has the following expression:
Dν(t)(cid:105) =
ην · Dj(t)ϕj(t)(cid:105) .
(C29)
j
In representing the final expressions, it is useful to intro-
duce the following polarization operators:
Pgν(t) = g(cid:105)(cid:104)Dν(t) ,
Pνg(t) = Dν(t)(cid:105)(cid:104)g = P †
gν(t) .
(C30)
(C31)
Expanding UT (t, t0) and U†
T (t, t0) with respect to
Hint(t), and collecting all the terms of the third order, we
find the following third order components of the density
operator:
†
†
ρ(3)(tm) = ρI (tm) + ρ
I (tm) + ρII (tm) + ρ
II (tm) , (C32)
(cid:88)
χ(2)(tm, t, t(cid:48), t(cid:48)(cid:48)) = T r
(cid:110) Ug(tm, t) Pg3(t) Uex(t, t(cid:48)(cid:48))
47
× P1g(t(cid:48)(cid:48)) Ug(t(cid:48)(cid:48), t0)ρ(t0) U†
× Pg2(t(cid:48)) U†
(cid:111)
g (t(cid:48), t0)
ex(tm, t(cid:48)) Pmg(tm)
,
(C48)
(cid:110) Ug(tm, t) Pg3(t) Uex(t, t(cid:48))
χ(3)(tm, t, t(cid:48), t(cid:48)(cid:48)) = T r
× P2g(t(cid:48)) Ug(t(cid:48), t0)ρ(t0) U†
× Pg1(t(cid:48)(cid:48)) U†
,
(cid:111)
g (t(cid:48)(cid:48), t0)
ex(tm, t(cid:48)(cid:48)) Pmg(tm)
(cid:110) Ug(tm, t0)ρ(t0) U†
(cid:111)
ex(t(cid:48), t(cid:48)(cid:48)) P2g(t(cid:48)) U†
g (t, t(cid:48))
ex(tm, t) Pmg(tm)
.
χ(4)(tm, t, t(cid:48), t(cid:48)(cid:48)) = T r
× Pg1(t(cid:48)(cid:48)) U†
× Pg3(t) U†
g (t(cid:48)(cid:48), t0)
(C49)
(C50)
The above general expressions for the response functions
are equivalent to those in previous works (Cho, 2008;
Cho et al., 2005; Jonas, 2003; Mukamel, 1995; Schlau-
Cohen et al., 2011a) within the assumption that only
single exciton states contribute. Fourier transforms of
these with respect to t(cid:48) − t(cid:48)(cid:48) and tm − t can be related to
the spectra of 2DES if proper averaging over the ensemble
of disorder is made.
Now, for the initial state given by Eq. (C11), the re-
sponse functions can be simplified as follows:
i(cid:82) t(cid:48)
t(cid:48)(cid:48) dτ Eg(τ )e− i(cid:82) tm
t
e− i (tm−t(cid:48)) HbDm(tm)(cid:105)(cid:104)D2(t(cid:48)) Uex(t(cid:48), t(cid:48)(cid:48))
i (t−t(cid:48)(cid:48)) Hb U†
ex(tm, t)
,
(C51)
(cid:110)
t
(cid:110)
χ(1)(tm, t, t(cid:48), t(cid:48)(cid:48)) = e
×T r
× ρbD1(t(cid:48)(cid:48))(cid:105)(cid:104)D3(t)e
χ(2)(tm, t, t(cid:48), t(cid:48)(cid:48)) = e− i(cid:82) tm
×T r
× ρbD1(t(cid:48)(cid:48))(cid:105)(cid:104)D2(t(cid:48))e
χ(3)(tm, t, t(cid:48), t(cid:48)(cid:48)) = e− i(cid:82) tm
×T r
×ρbD2(t(cid:48))(cid:105)(cid:104)D1(t(cid:48)(cid:48))e
χ(4)(tm, t, t(cid:48), t(cid:48)(cid:48)) = e− i(cid:82) tm
×T r
×D2(t(cid:48))(cid:105)(cid:104)D3(t)e
(cid:110)
(cid:110)
t
i (t(cid:48)−t(cid:48)(cid:48)) Hb U†
ex(tm, t(cid:48))
dτ Eg(τ )e− i(cid:82) t(cid:48)
t(cid:48)(cid:48) dτ Eg(τ )
,(C52)
(cid:111)
e− i (tm−t) HbDm(tm)(cid:105)(cid:104)D3(t) Uex(t, t(cid:48))
ex(tm, t(cid:48)(cid:48))
i (t(cid:48)−t(cid:48)(cid:48)) Hb U†
dτ Eg(τ )e− i(cid:82) t(cid:48)
(cid:111)
e− i (tm−t(cid:48)(cid:48)) HbDm(tm)(cid:105)(cid:104)D1(t(cid:48)(cid:48))ρb U†
t(cid:48)(cid:48) dτ Eg(τ )
t
i (t−t(cid:48)) Hb U†
ex(tm, t)
.
ex(t(cid:48), t(cid:48)(cid:48))
(C54)
,(C53)
dτ Eg(τ )
(cid:111)
(cid:111)
i(cid:82) t(cid:48)
t(cid:48)(cid:48) dτ Eg(τ )
e− i (tm−t) HbDm(tm)(cid:105)(cid:104)D3(t) Uex(t, t(cid:48)(cid:48))
dτ Eg(τ )e
0
0
0
0
0
0
dt
= e−i(k3+k2−k1)·rei(ω3t3+ω2t2−ω1t1)
×
dt(cid:48)(cid:48) ηm · T r
(cid:90) tm
(cid:90) t
(cid:90) tm
dt(cid:48)(cid:90) t(cid:48)
dt(cid:48)(cid:90) t(cid:48)
(cid:90) t
dt
3 (t − t3)eiω3(t−t3)E∗
×E∗
×E1(t(cid:48)(cid:48) − t1)e−iω(t(cid:48)(cid:48)−t1)
(cid:90) t
(cid:90) tm
(cid:90) tm
dt(cid:48)(cid:90) t(cid:48)
dt(cid:48)(cid:90) t(cid:48)
(cid:90) t
dt
3 (t − t3)eiω3(t−t3)E∗
×E∗
×E1(t(cid:48)(cid:48) − t1)e−iω1(t(cid:48)(cid:48)−t1) ,
(cid:90) tm
(cid:90) t
(cid:90) tm
dt(cid:48)(cid:48) ηm · T r
dt(cid:48)(cid:48) ηm · T r
= e−i(k3+k2−k1)·rei(ω3t3+ω2t2−ω1t1)
×
dt
0
0
0
0
0
0
0
0
(C44)
(C45)
0
0
0
0
dt
dt(cid:48)(cid:48)χ(3)(tm, t, t(cid:48), t(cid:48)(cid:48))
(cid:111)
(cid:110) P T3(tm, t, t(cid:48), t(cid:48)(cid:48))
= e−i(k3−k2+k1)·rei(ω3t3−ω2t2+ω1t1)
×
dt(cid:48)(cid:90) t(cid:48)
(cid:90) t
dt(cid:48)(cid:90) t(cid:48)
dt
3 (t − t3)eiω3(t−t3)E2(t(cid:48) − t2)e−iω2(t(cid:48)−t2)
×E∗
1 (t(cid:48)(cid:48) − t1)eiω1(t(cid:48)(cid:48)−t1) ,
×E∗
(cid:90) tm
(cid:90) t
(cid:111)
(cid:110) P T4(tm, t, t(cid:48), t(cid:48)(cid:48))
(cid:90) tm
dt(cid:48)(cid:90) t(cid:48)
dt(cid:48)(cid:90) t(cid:48)
(cid:90) t
dt
×E∗
3 (t − t3)eiω3(t−t3)E2(t(cid:48) − t2)e−iω2(t(cid:48)−t2)
1 (t(cid:48)(cid:48) − t1)eiω1(t(cid:48)(cid:48)−t1) .
×E∗
= e−i(k3−k2+k1)·rei(ω3t3−ω2t2+ω1t1)
×
dt(cid:48)(cid:48)χ(4)(tm, t, t(cid:48), t(cid:48)(cid:48))
dt(cid:48)(cid:48) ηm · T r
dt
0
0
0
0
0
0
(C46)
where χ(k), k = 1 − 4, are response functions. Under the
assumption that excited state absorption or double exci-
ton formation can be neglected, these can be expressed
as
(cid:110) Ug(tm, t(cid:48)) Pg2(t(cid:48)) Uex(t(cid:48), t(cid:48)(cid:48))
χ(1)(tm, t, t(cid:48), t(cid:48)(cid:48)) = T r
× P1g(t(cid:48)(cid:48)) Ug(t(cid:48)(cid:48), t0)ρ(t0) U†
× Pg3(t) U†
ex(tm, t) Pmg(tm)
g (t, t0)
(cid:111)
Then, taking scalar product of ηm with the integrand of
Eq. (C40) and considering only those terms where inter-
actions with E1, E2, and E3 occur in the chronological
order at t(cid:48)(cid:48), t(cid:48), and t, respectively, we obtain the following
general expressions:
(cid:110) P T1(tm, t, t(cid:48), t(cid:48)(cid:48))
(cid:111)
dt(cid:48)(cid:48)χ(1)(tm, t, t(cid:48), t(cid:48)(cid:48))
2 (t(cid:48) − t2)eiω2(t(cid:48)−t2)
(C43)
(cid:110) P T2(tm, t, t(cid:48), t(cid:48)(cid:48))
(cid:111)
dt(cid:48)(cid:48)χ(2)(tm, t, t(cid:48), t(cid:48)(cid:48))
2 (t(cid:48) − t2)eiω2(t(cid:48)−t2)
,
(C47)
For the case where the Hamiltonians are time indepen-
dent, the above expressions become Eqs. (51)-(54).
48
REFERENCES
Abramavicius, D., L. Valkunas, and R. van Grondelle (2004),
Phys. Chem. Chem. Phys. 6, 3097.
Abramavicius, D., and S. Mukamel (2011), J. Chem. Phys.
134, 174504.
Brixner, T., J. Stenger, H. M. Vaswani, M. Cho, R. E.
Blankenship, and G. R. Fleming (2005), Nature 434, 625.
Brotosudamo, T. H. P., R. Kunz, P. Bohm, A. T. Gardiner,
V. Moulisov´a, R. J. Cogdell, and J. Kohler (2009), Bio-
phys. J. 97, 1491.
Bruggemann, B., J. L. Herek, V. Sundstrom, and V. May
Abramavicius, D., B. Palmieri, D. V. Voronine, F. Sanda,
(2001), J. Phys. Chem. B 105, 11391.
and S. Mukamel (2009), Chem. Rev. 109, 2350.
Bruggemann, B., and V. May (2004), J. Phys. Chem. B 108,
Abramavicius, V., and D. Abramavicius (2014), J. Chem.
10529.
Phys. 140, 065103.
Brumer, P., and M. Shapiro (2012), Proc. Natl. Acad. Sci.,
Adolphs, J., F. Mueh, M. E.-A. Madjet,
and T. Renger
USA 109, 19575.
(2008), Photosyn. Res. 95, 197.
Adolphs, J., and T. Renger (2006), Biophys. J. 91, 2778.
Agarwal, R., A. H. Rizvi, B. S. Prall, J. D. Olsen, C. N.
Hunter, and G. R. Fleming (2002), J. Phys. Chem. A 106,
7573.
Aghtar, M., U. Kleinekathofer, C. Curutchet, and B. Men-
nucci (2017), J. Phys. Chem. B 121, 1330.
Aghtar, M., J. Strumpfer, C. Olbrich, K. Schulten,
and
U. Kleinekathofer (2013), J. Phys. Chem. B 117, 7157.
Aghtar, M., J. Strumpfer, C. Olbrich, K. Schulten,
and
U. Kleinekathofer (2014), J. Phys. Chem. Lett. 5, 3131.
Agranovich, V. M., and M. D. Galanin (1982), Modern Prob-
lems in Condensed Matter Sciences: Vol.3, Electronic exci-
tation energy transfer in condensed matter (North-Holland,
Amsterdam).
Agranovich, V. M., and R. M. Hochstrasser, Eds. (1982),
Modern Problems in Condensed Matter Sciences: Vol.4,
Spectroscopy and excitation dynamics of condensed molec-
ular systems (North-Holland, Amsterdam).
Alden, R., E. Johnson, V. Nagarajan, W. W. Parson, C. J.
Law, and R. J. Cogdell (1997), J. Phys. Chem. B 101,
4667.
van Amerongen, H., R. van Grondelle,
and L. Valkunas
(2000), Photosynthetic Excitons (World Scientific, River
Edge, NJ).
van Amerongen, H., and W. S. Struve (1991), J. Phys. Chem.
95, 9020.
Buck, D. R., S. Savikhin, and W. S. Struve (1997), Biophys.
J. 72, 24.
Butkus, V., D. Zigmantas, L. Valkunas, and D. Abramavicius
(2012), Chem. Phys. Lett. 545, 40.
Cao, J., and R. J. Silbey (2009), J. Phys. Chem. A 113,
13825.
Caprasecca, S., C. Curutchet, and B. Mennucci (2012), J.
Chem. Theory Comput. 8 (11), 4462.
Carbonera, D., G. Giacometti, U. Segre, E. Hofmann, and
R. G. Hiller (1999), J. Phys. Chem. B 103 (30), 6349.
Cartron, M. L., J. D. Olsen, M. Sener, P. J. Jackson, A. A.
Brindley, P. Qian, M. J. Dickman, G. J. Leggett, K. Schul-
ten, and C. N. Hunter (2014), Biochim. Biophys. Acta
1837, 1769.
Caruso, F., A. W. Chin, A. Datta, S. F. Huelga, and M. B.
Plenio (2009), J. Chem. Phys. 131, 105106.
Caruso, F., A. W. Chin, A. Datta, S. F. Huelga, and M. B.
Plenio (2010), Phys. Rev. A 81, 062346.
Caycedo-Soler, F., F. J. Rodriguez, L. Quiroga, and N. F.
Johnson (2010), Phys. Rev. Lett. 104, 158302.
Chandrasekaran, S., M. Aghtar, S. Valleau, A. Aspuru-Guzik,
and U. Kleinekathofer (2015), J. Phys. Chem. B 10, 9995.
Chandrasekaran, S., K. R. Pothula, and U. Kleinekathofer
(2017), J. Phys. Chem. B 121, 3228.
Chang, J. C. (1977), J. Chem. Phys. 67 (9), 3901.
Chen, G.-Y., N. Lambert, C.-M. Li, Y.-N. Chen, and F. Nori
(2013), Phys. Rev. E 88, 032120.
Anda, A., T. Hansen, and L. D. Vico (2016), J. Chem. Theory
Chen, L., R. Zheng, Q. Shi, and Y. Yan (2009), J. Chem.
Comput. 12, 1305.
Andreussi, O., S. Knecht, C. M. Marian, J. Kongsted, and
B. Mennucci (2015), J. Chem. Theory Comput. 11 (2), 655.
Bahatyrova, S., R. N. Frese, C. A. Siebert, J. D. Olsen, K. O.
van der Werf, R. van Grondelle, R. A. Niederman, P. A.
Bullough, C. Otto, and N. Hunter (2004), Nature 430,
1058.
Berlin, Y., A. Burin, J. Friedrich, and J. Kohler (2007), Phys.
Life. Rev. 4, 64.
Phys. 131, 094502.
Cheng, Y. C., and G. R. Fleming (2009), Annu. Rev. Phys.
Chem. 60, 241.
Cheng, Y. C., and R. J. Silbey (2006), Phys. Rev. Lett. 96,
028103.
Chenu, A., A. M. Branczyk, and G. D. Scholes (2015), Phys.
Rev. Lett. 114, 213601.
Chenu, A.,
and P. Brumer (2016), J. Chem. Phys. 144,
044103.
Blankenship, R. E. (2014), Molecular Mechanisms of Photo-
Chenu, A., P. Mal'y, and T. Mancal (2014), Chem. Phys.
synthesis, 2nd Ed. (Wiley Blackwell, Oxford).
Book, L. D., A. E. Ostafin, N. Ponomarenko, J. R. Norris,
and N. F. Scherer (2000), J. Phys. Chem. B 104, 8295.
Bopp, M. A., Y. Jia, L. Li, R. J. Cogdell,
and R. M.
Hochstrasser (1997), Proc. Natl. Acad. Sci. 94, 10630.
Bopp, M. A., A. Sytnik, T. D. Howard, R. J. Cogdell, and
R. M. Hochstrasser (1999), Proc. Natl. Acad. Sci. 96,
11271.
Boulanger, E., and W. Thiel (2012), J. Chem. Theory Com-
439, 100.
Chenu, A.,
and G. D. Scholes (2015), Annu. Rev. Phys.
Chem. 66, 69.
Chin, A. W., A. Datta, F. Caruso, S. F. Huelga, and M. B.
Plenio (2010), New J. Phys. 12, 065002.
Chin, A. W., J. Prior, R. Rosenbach, F. Caycedo-Soler, S. F.
Huelga, and M. B. Plenio (2013), Nature Phys. 9, 113.
Cho, M. (2008), Chem. Rev. 108, 1331.
Cho, M., H. M. Vaswani, T. Brixner, J. Stenger, and G. R.
Fleming (2005), J. Phys. Chem. B 109, 10542.
put. 8, 4527.
Bricker, W. P.,
118 (31), 9141.
and C. S. Lo (2014), J. Phys. Chem. B
Christensson, N., H. Kauffmann, T. Pullerits, and T. Mancal
(2012), J. Phys. Chem. B 116, 7449.
Bricker, W. P.,
and C. S. Lo (2015), J. Phys. Chem. B
Cleary, L., H. Chen, C. Chuang, R. J. Silbey, and J. Cao
119 (18), 5755.
(2013), Proc. Natl. Acad. Sci. USA 110, 8537.
Cogdell, R. J., A. Gali, and J. Kohler (2006), Quart. Rev.
49
Biophys. 39, 227.
Cohen, G., E. Y. Wilner, and E. Rabani (2013), New J. Phys.
15, 073018.
Cole, D. J., A. W. Chin, N. D. M. Hine, P. D. Haynes, and
M. C. Payne (2013), J. Phys. Chem. Lett. 4, 4206.
Fenna, R. E., and B. W. Matthews (1975), Nature 258, 573.
Ferretti, M., R. Hendrikx, E. Romero, J. Southall, R. J.
Cogdell, V. I. Novoderezhkin, G. D. Scholes, and R. van
Grondelle (2016), Sci. Rep. 6, 20834.
Fleming, C. H., and N. I. Cummings (2011), Phys. Rev. E
Collini, E., C. Y. Wong, K. E. Wilk, P. M. G. Curmi,
83, 03117.
P. Brumer, and G. D. Scholes (2010), Nature 463, 644.
Cotton, S. J.,
and W. H. Miller (2016), J. Chem. Theoy
Comput. 12, 983.
Croce, R., and H. van Amerongen (2014), Nat. Chem. Bio.
10, 492.
Cupellini, L., S. Jurinovich, M. Campetella, S. Caprasecca,
C. A. Guido, S. M. Kelly, A. T. Gardiner, R. Cogdell, and
B. Mennucci (2016), J. Phys. Chem. B 120, 11348.
Curutchet, C., J. Kongsted, A. M. noz Losa, H. Hossein-
Nejad, G. D. Scholes, and B. Mennucci (2011a), J. Am.
Chem. Soc. 133, 3078.
Curutchet, C., J. Kongsted, A. Munoz-Losa, H. Hossein-
Nejad, G. D. Scholes, and B. Mennucci (2011b), Journal
of the American Chemical Society 133 (9), 3078.
Curutchet, C., A. M. noz Losa, S. Monti, J. Kongsted, G. D.
Scholes, and B. Mennucci (2009), J. Chem. Theory Com-
put. 5, 18238.
Fleming, G. R., and G. D. Scholes (2004), Nature 431, 256.
Forster, T. (1948), Ann. Phys. 6, 55.
Forster, T. (1959), Discuss. Faraday Soc. 27, 7.
Fransted, K. A., J. Caram, D. Hayes, and G. S. Engel (2012),
J. Chem. Phys. 137, 125101.
Freiberg, A., S. Lin, K. Timpmann, and R. E. Blankenship
(1997), J. Phys. Chem. B 101, 7211.
Frenkel, J. (1931), Phys. Rev. 37, 17.
Fujihashi, Y., G. R. Fleming,
and A. Ishizaki (2015), J.
Chem. Phys. 142, 212403.
Fuller, F. D., and J. P. Ogilvie (2015), Annu. Rev. Phys.
Chem. 66, 667.
Georgakopoulou, S., R. N. Frese, E. Johnson, C. Koolhaas,
R. J. Cogdell, R. van Grondelle, and G. van der Zwan
(2002), Biophys. J. 82, 2184.
Ghosh, P. K., A. Yu, and F. Nori (2009), J. Chem. Phys.
131, 035102.
Curutchet, C., and B. Mennucci (2017), Chem. Rev. 113,
Ginsberg, N. S., Y.-C. Cheng, and G. R. Fleming (2009),
294.
Curutchet, C., V. I. Novoderezhkin, J. Kongsted, A. M. noz
Losa, R. van Grondelle, G. D. Scholes, and B. Mennucci
(2013), J. Phys. Chem. B 117, 4283.
Acc. Chem. Res. 42, 1352.
Grimme, S., and M. Waletzke (1999), J. Comp. Phys. 111,
5645.
Grinev, T.,
and P. Brumer (2015), J. Chem. Phys. 143,
Curutchet, C., G. D. Scholes, B. Mennucci, and R. Cammi
244313.
(2007), J. Phys. Chem. B 111, 13253.
Dahlbom, M., T. Pullerits, S. Mukamel, and V. Sundstrom
(2001), J. Phys. Chem. B 105, 5515.
Gulen, D. (1996), J. Phys. Chem. 100, 17683.
Haken, H., and G. Strobl (1973), Z. Phys. 262, 5.
Harel, E., and G. Engel (2012), Proc. Natl. Acad. Sci. USA
Damjanovi´c, A., I. Kosztin, U. Kleinekathofer, and K. Schul-
109, 706.
ten (2002), Phys. Rev. E 65, 031919.
Davydov, A. S. (1971), Theory of molecular excitons (Plenum
press, NewYork-London).
Dean, J., T. Mirkovic, Z. S. D. Toa, D. G. Oblinsky, and
G. D. Scholes (2016), Chem. 1, 858.
Dempster, S. E., S. Jang, and R. J. Silbey (2001), J. Chem.
Phys. 114, 10015.
Dinh, T.-C., and T. Renger (2016), J. Chem. Phys. 145,
034105.
Harrop, S. J., K. E. Wilk, R. Dinshaw, E. Collini, T. Mirkovic,
C. Y. Teng, D. G. Oblinsky, B. R. Green, K. Hoef-Emden,
R. G. Hiller, G. D. Scholes, and P. M. G. Curmi (2014),
Proc. Natl. Acad. Sci., USA 111, E2666.
Hayes, D., and G. S. Engel (2011), Biophys. J. 100, 2043.
Hayes, D., G. Panitchayangkoon, K. A. Fransted, J. R.
Caram, J. Wen, K. F. Freed, and G. S. Engel (2010), New.
J. Phys. 12, 065042.
Hein, B., C. Kreisbeck, T. Kramer, and M. Rodriguez (2012),
Dost´al, J., J. Psenc´ık, and D. Zigmantas (2016), Nat. Chem.
New. J. Phys. 14, 023018.
8, 705.
Doust, A. B., C. N. J. Marai, S. J. Harrop, K. E. Wilk,
P. M. G. Curmi, and G. D. Scholes (2004), J. Mol. Biol.
344, 135.
Doust, A. B., I. H. M. van Stokkum, D. S. Larsen, K. E.
Wilk, P. M. G. Curmi, R. van Grondelle, and G. D. Scholes
(2005), J. Phys. Chem. B 109, 14219.
Duan, H.-G., V. I. Prokhorenko, R. J. Cogdell, K. Ashraf,
and R. J. Dwayne-Miller
A. L. Stevens, M. Thorwart,
(2017), Proc. Natl. Acad. Sci., USA 114, 8493.
Duffy, C. D. P., J. Chmeliov, M. Macernis, J. Sulskus,
and a. V. Ruban (2013), J. Phys. Chem.
L. Valkunas,
B 117 (38), 10974.
Engel, G. S., T. R. Calhoun, E. L. Read, T.-K. Ahn, T. Man-
cal, Y.-C. Cheng, R. E. Blankenship, and G. R. Fleming
(2007), Nature 446, 782.
Feng, J., C.-W. Tseng, T. Chen, X. Leng, H. Yin, Y.-C.
Cheng, M. Rohlfing, and Y. Ma (2017), Nat. Commun.
8, 71.
Herek, J. L., N. J. Fraser, T. Pullerits, P. Martinsson,
T. Pol´ıvka, H. Scheer, R. J. Cogdell, and V. Sundstrom
(2000), Biophys. J. 78, 2590.
Hildner, R., D. Brinks, J. B. Nieder, R. J. Cogdell, and N. F.
van Hulst (2013), Science 340, 1448.
Hofmann, C., T. J. Aartsma, and J. Kohler (2004), Chem.
Phys. Lett. 395, 373.
Hossein-Nejad, H., C. Curutchet, A. Kubica, and G. D. Sc-
holes (2011), J. Phys. Chem. B 115, 5243.
Hoyer, S., M. Sarovar, and K. B. Whaley (2010), New. J.
Phys. 12, 065041.
Hsu, C.-P. (2009), Acc Chem Res 42 (4), 509.
Hsu, C.-P., G. R. Fleming, M. Head-Gordon, and T. Head-
Gordon (2001a), J. Chem. Phys. 114 (7), 3065.
Hsu, C.-P., P. J. Walla, M. Head-Gordon, and G. R. Fleming
(2001b), J. Phys. Chem. B 105, 11016.
Hsu, C.-P., Z.-Q. You, and H.-C. Chen (2008), J. Phys. Chem.
C 112 (4), 1204.
Hu, X., T. Ritz, A. Damjanovic, F. Autenrieth, and K. Schul-
Fenna, R. E., L. F. T. Eyck, and B. W. Matthews (1977),
ten (2002), Quar. Rev. Biophys. 35, 1.
Biochem. Biophys. Res. Commun. 75, 751.
Hu, X., T. Ritz, A. Damjanovic, and K. Schulten (1997), J.
50
Phys. Chem. B 101, 3854.
Kassal, I., J. Yuen-Zhou, and S. Rahimi-Keshari (2013), J.
Huh, J., S. K. Saikin, J. C. Brookes, S. Valleau, T. Fujita, and
Phys. Chem. Lett. 4, 362.
A. Aspuru-Guzik (2013), J. Am. Chem. Soc. 136, 2048.
Huo, P., and D. F. Coker (2010), J. Chem. Phys. 133, 184108.
Huo, P., and D. F. Coker (2011), J. Phys. Chem. Lett. 2,
Kell, A., K. Acharya, V. Zaaubovich,
(2014), J. Phys. Chem. Lett. 5, 1450.
and R. Jankowiak
Kell, A., X. Feng, M. Reppert, and R. Jankowiak (2013), J.
825.
Phys. Chem. B 117, 7317.
Hwang-Fu, Y. H., W. Chen, and Y.-C. Cheng (2015), Chem.
Kelly, A., A. Montoya-Castillo, L. Wang, and T. E. Markland
Phys. 447, 46.
(2016), J. Chem. Phys. 144, 184105.
Iles-Smith, J., A. G. Dijkstra, N. Lambert, and A. Nazir
Kelly, A., and Y. M. Rhee (2011), J. Phys. Chem. Lett. 2,
(2016), J. Chem. Phys. 144, 044110.
808.
Iles-Smith, J., N. Lambert, and A. Nazir (2014), Phys. Rev.
A 90, 032114.
Kenkre, V. M., and R. S. Knox (1974), Phys. Rev. B 9, 5279.
Kenkre, V. M., and P. Reineker (1982), Exciton Dynamics
Iozzi, M. F., B. Mennucci, J. Tomasi, and R. Cammi (2004),
in Molecular Crystals and Aggregates (Springer, Berlin).
J. Chem. Phys. 120 (15), 7029.
Kenny, E. P., and I. Kassal (2016), J. Phys. Chem. B 120 (1),
Ishizaki, A., and G. R. Fleming (2009a), Proc. Natl. Acad.
25.
Sci., USA 106, 17255.
Ishizaki, A., and G. R. Fleming (2009b), J. Chem. Phys. 130,
234111.
Ketelaars, M., A. M. van Oijen, M. Matsushita, J. Kohler,
J. Schmidt, and T. J. Aartsma (2001), Biophys. J. 80,
1591.
Ishizaki, A., and G. R. Fleming (2011), J. Phys. Chem. B
Killoran, N., S. F. Huelga, and M. B. Plenio (2015), J. Chem.
115, 6227.
Phys. 143, 155102.
J. Strumpfer and K. Schulten, (2012), J. Chem. Phys. 137,
Kim, C. W., B. Choi, and Y. M. Rhee (2018), Phys. Chem.
065101.
Jang, S. (2007), J. Chem. Phys. 127, 174710.
Jang, S. (2011), J. Chem. Phys. 135, 034105.
Jang, S. (2012), Bull. Kor. Chem. Soc. 33, 997.
Jang, S., T. Berkelbach, and D. R. Reichman (2013), New J.
Chem. Phys. 20, 3310.
Kim, C. W., J. W. Park, and Y. M. Rhee (2015), J. Phys.
Chem. Lett. 6, 2875.
Kim, C. W., and Y. M. Rhee (2016), J. Chem. Theory Com-
put. 12, 5235.
Phys. 15, 105020.
Kimura, A., and T. Kakitani (2003), J. Phys. Chem. B ,
Jang, S., J. Cao, and R. J. Silbey (2002a), J. Chem. Phys.
7932.
116, 2705.
Klamt, A. (2011), Wiley Interdisciplinary Reviews: Compu-
Jang, S., and Y.-C. Cheng (2013), WIREs Comput. Mol. Sci.
tational Molecular Science 1 (5), 699.
3, 84.
Kleinekathofer, M. S. U., and M. Schreiber (2006), J. Chem.
Jang, S., S. E. Dempster, and R. J. Silbey (2001), J. Phys.
Phys. 124, 084903.
Chem. B 105, 6655.
Jang, S., S. Hoyer, G. R. Fleming, and K. B. Whaley (2014),
Phys. Rev. Lett. 113, 188102.
Jang, S., Y. J. Jung, and R. J. Silbey (2002b), Chem. Phys.
275, 319.
Kleinschmidt, M., C. M. Marian, M. Waletzke,
S. Grimme (2009), J. Comp. Phys. 130 (4), 044708.
and
Knox, R. S. (1996), Photosynth. Res. 48, 35.
Koepke, J., X. Hu, C. Muenke, K. Schulten, and H. Michel
(1996), Structure 4, 581.
Jang, S., M. D. Newton, and R. J. Silbey (2004), Phys. Rev.
Kohen, D., C. C. Marston, and D. J. Tannor (1997), J. Chem.
Lett. 92, 218301.
Phys. 107, 5236.
Jang, S., M. D. Newton, and R. J. Silbey (2007), J. Phys.
Kolli, A., E. J. O'Reilly, G. D. Scholes, and A. Olaya-Castro
Chem. B 111, 6807.
Jang, S., E. Rivera, and D. Montemayor (2015), J. Phys.
Chem. Lett. 6, 928.
Jang, S., and R. J. Silbey (2003a), J. Chem. Phys. 118, 9324.
Jang, S., and R. J. Silbey (2003b), J. Chem. Phys. 118, 9312.
Jang, S., R. J. Silbey, R. Kunz, C. Hofmann, and J. Kohler
(2012), J. Chem. Phys. 137, 174109.
Kondo, T., W. J. Chen,
Chem. Rev. 117, 860.
and G. S. Schlau-Cohen (2017),
Konig, C., and J. Neugebauer (2012), Chem. Phys. Chem.
13, 386.
Kreisbeck, C., and T. Kramer (2012), J. Phys. Chem. Lett.
(2011), J. Phys. Chem. B 115, 12947.
3, 2828.
Jankowiak, R., M. Peppert, V. Zazubovich, J. Pieper, and
Krueger, B. P., G. D. Scholes, and G. R. Fleming (1998), J.
T. Reinot (2011), Chem. Rev. 111, 4546.
Phys. Chem. B 102, 5378.
Jansen, T. L. C., and J. Knoester (2009), Acc. Chem. Res.
42, 1405.
Jimenez, R., S. Dikshit, S. E. Bradforth, and G. R. Fleming
(1996), J. Phys. Chem. 100, 6825.
Jimenez, R., F. van Mourik, J. Y. Yu, and G. R. Fleming
(1997), J. Phys. Chem. B 101, 7350.
Jing, Y., L. Chen, S. Bai, and Q. Shi (2013), J. Chem. Phys.
138, 045101.
Johnson, S. G., and G. J. Small (1991), J. Phys. Chem. 95,
471.
Jonas, D. M. (2003), Annu. Rev. Phys. Chem. 54, 425.
Jurinovich, S., C. Curutchet,
and B. Mennucci (2014),
Kumar, P., and S. Jang (2013), J. Chem. Phys. 138, 135101.
Kunz, R., K. Timpmann, J. Southall, R. J. Cogdell,
A. Freiberg, and J. Kohler (2012), J. Phys. Chem. B 116,
11017.
Kunz, R., K. Timpmann, J. Southall, R. J. Cogdell,
A. Freiberg, and J. Kohler (2014), Biophys. J. 106, 2008.
Lambert, N., Y.-N. Chen, Y.-C. Cheng, C.-M. Li, and G.-Y.
Chen (2013), Nature Phys. 9, 10.
Lampoura, S., R. van Grondelle, I. H. M. van Stokkum, R. J.
Cogdell, D. A. Wiersma, and K. Duppen (2000), J. Phys.
Chem. B 104, 12072.
Lee, M. K., K. B. Bravaya, and D. F. Coker (2017), J. Am.
ChemPhysChem 15 (15), 3194.
Chem. Soc. 139, 7803.
Kananeka, A. A., C. Y. Hsieh, J. Cao, and E. Geva (2016),
Lee, M. K., and D. F. Coker (2016), J. Phys. Chem. Lett. 7,
J. Phys. Chem. Lett. 7, 4809.
3171.
51
Lee, M. K., P. Huo, and D. F. Coker (2016), Annu. Rev.
Phys. Chem. 67, 639.
Leegwater, J. A. (1996), J. Phys. Chem. 100, 14403.
Levi, F., S. Mostarda, F. Rao, and F. Mintert (2015), Rep.
B. Rabenstein, H. Ishikita, E.-W. Knapp, and T. Renger
(2007), Proc. Natl. Acad. Sci., USA 104, 16862.
Muhlbacher, L.,
and U. Kleinekathofer (2012), J. Phys.
Chem. B 116, 3900.
Prog. Phys. 78, 082001.
Mukai, K., S. Abe, and H. Sumi (1999), J. Phys. Chem. B
Lindblad, G. (1976), Commun. Math. Phys. 48, 119.
Lipparini, F., and V. Barone (2011), J. Chem. Theory Com-
put. 7 (11), 3711.
Liu, X., and O. Kuhn (2016), Chem. Phys. 481, 272.
Liuolia, V., L. Valkunas, and R. van Grondelle (1997), J.
Phys. Chem. B 101, 7347.
Lohner, A., K. Ashraf, R. J. Cogdell, and J. Kohler (2016),
Sci. Rep. 6, 31875.
Lohner, A., A.-M. Carey, K. Hacking, N. Picken, S. Kelly,
R. Cogdell, and J. Kohler (2015), Photosyn. Res. 123, 23.
Louwe, R. J. W., J. Vrieze, A. J. Hoff, and T. J. Aartsma
103, 6096.
Mukamel, S. (1995), Principles of Nonlinear Spectroscopy
(Oxford University Press, New York).
Munoz-Losa, A., C. Curutchet, I. F. Galv´an, and B. Men-
nucci (2008), J. Chem. Phys. 129 (3), 034104.
Nazir, A. (2009), Phys. Rev. Lett. 103, 146404.
Novoderezhkin, V. I., A. B. Doust, C. Curutchet, G. D. Sc-
holes, and R. van Grondelle (2010), Biophys. J. 99, 344.
Novoderezhkin, V. I., and R. van Grondelle (2017), J. Phys.
B: Mol. Opt. Phys. 50, 124003.
Novoderezhkin, V. I., M. Wendling, and R. van Grondelle
(1997), J. Phys. Chem. B 101, 11280.
(2003), J. Phys. Chem. B 107, 11534.
Ma, J., and J. Cao (2015), J. Chem. Phys. 142, 094106.
Madjet, M. E., A. Abdurahman, and T. Renger (2006), J.
Oellerich, S., and J. Kohler (2009), Photosynth. Res. 101,
171.
Phys. Chem. B 110, 17268.
van Oijen, A. M., M. Ketelaars, J. Kohler, T. J. Aartsma,
Madjet, M. E.-A., F. Muh, and T. Renger (2009), J. Phys.
and J. Schmidt (1999), Science 285, 400.
Chem. B 113 (37), 12603.
Magdaong, N. C. M., R. G. Saer, D. M. Niedzwiedzki, and
R. E. Blankenship (2017), J. Phys. Chem. B 121, 4700.
McClure, S. D., D. B. Turner, P. C. Arpin, T. Mirkovic, and
G. D. Scholes (2014), J. Phys. Chem. B 118, 1296.
Olaya-Castro, A., A. Nazir, and G. R. Fleming, Eds. (2012),
Phil. Trans. A, Vol. 370, no. 1972 - Quantum-coherent
energy transfer:
implications for biology and new energy
transfer (Royal Society Publishing, London).
Olaya-Castro, A., and G. D. Scholes (2011), Int. Rev. Phys.
McCutcheon, D. P. S., and A. Nazir (2011), J. Chem. Phys.
Chem. 30, 49.
135, 114501.
Olbrich, C., and U. Kleinekathofer (2010), J. Phys. Chem. B
McDermott, G., S. M. Prince, A. A. Freer, A. M.
Hawthornthwaite-Lawless, M. Z. Paplz, R. J. Cogdell, and
N. W. Issacs (1995), Nature 374, 517.
114, 12427.
Olbrich, C., J. Strumpfer, K. Schulten, and U. Kleinekathofer
(2011a), J. Phys. Chem. B 115, 8609.
McRobbie, P. L., and E. Geva (2009), J. Phys. Chem. A 113,
Olbrich, C., J. Strumpfer, K. Schulten, and U. Kleinekathofer
10425.
(2011b), J. Phys. Chem. Lett. 2, 1771.
Meier, T., V. Chernyak, and S. Mukamel (1997a), J. Chem.
Olsen, J. D., J. D. Tucker, J. A. Timney, P. Qian, C. Vassilev,
Phys. 107, 8759.
and C. N. Hunter (2008), J. Biol. Chem. 283, 30772.
Meier, T., V. Chernyak, and S. Mukamel (1997b), J. Phys.
O'Reilly, E. J., and A. Olya-Castro (2014), Nat. Commun.
Chem. B 101, 7332.
5, 3012.
Meier, T., Y. Zhao, V. Chernyak, and S. Mukamel (1997c),
Ostroumov, E. E., R. Mulvaney, R. J. Cogdell, and G. D.
J. Chem. Phys. 107, 3876.
Scholes (2013), Science 340, 52.
Mennucci, B. (2012), WIREs Comput. Mol. Sci. 2, 386.
Mennucci, B. (2013), Phys. Chem. Chem. Phys. 15, 6583.
Mercer, I. P., Y. C. El-Taha, N. Kajumba, J. P. Marangos,
J. W. G. Tisch, M. Gabrielsen, R. J. Cogdell, E. Springate,
and E. Turcu (2009), Phys. Rev. Lett. 102, 057402.
Pach´on, L. A., and P. Brumer (2012), Phys. Chem. Chem.
Phys. 14, 10094.
Pajusalu, M., M. Ratsep, G. Trinkunas,
and A. Freiberg
(2011a), Chem. Phys. Chem. 12, 634.
Pajusalu, M., M. Ratsep, G. Trinkunas,
and A. Freiberg
Miller, M., R. P. Cox, and J. M. Olson (1994), Photosyn.
(2011b), ChemPhysChem 12 (3), 634.
Res. 41, 97.
Miller, W. H. (2001), J. Phys. Chem. A 105, 2942.
Mirkovic, T., A. B. Doust, J. Kim, E. Wilk, C. Curutchet,
B. Mennucci, R. Cammi, P. M. G. Curmi, and G. D. Sc-
holes (2007), Photochem. Photobiol. Sci. 6, 964.
Mirkovic, T., E. E. Ostroumov, J. M. Anna, R. van Grondelle,
Govindjee, and G. D. Scholes (2017), Chem. Rev. 113, 249.
Mohseni, M., P. Rebentrost, S. Lloyd, and A. Aspuru-Guzik
(2008), J. Chem. Phys. 129, 174106.
Palenberg, M. A., R. J. Silbey, C. Warns, and P. Reineker
(2001), J. Chem. Phys. 114, 4386.
Palmieri, B., D. Abramavicius, and S. Mukamel (2009), J.
Chem. Phys. 130, 204512.
Panitchayangkoon, G., D. Hayes, K. A. Fransted, J. R.
Caram, E. Harel, J. Wen, R. E. Blankenship, and G. S.
Engel (2010), Proc. Natl. Acad. Sci. USA 107, 12766.
Papiz, M. Z., S. M. Prince, T. Howard, R. J. Cogdell, and
N. W. Isaacs (2003), J. Mol. Biol. 326 (5), 1523.
Mohseni, M., A. Shabani, S. Lloyd, and H. Rabitz (2014), J.
Park, J. W., and Y. M. Rhee (2012), J. Phys. Chem. B 116,
Chem. Phys. 140, 035102.
11137.
Moix, J. M., J. Wu, P. Huo, D. Coker, and J. Cao (2011), J.
Phys. Chem. Lett. 2, 3045.
Montemayor, D., E. Rivera, and S. J. Jang (2018), J. Phys.
Chem. B 122, 3815.
Mostovoy, M. V., and J. Knoester (2000), J. Phys. Chem. B
104, 12355.
Mourokh, L. G., and F. Nori (2015), Phys. Rev. E 92, 052720.
Muh, F., M. E. Madjet, J. Adolphs, A. Abdurahman,
Pearlstein, R. M. (1992), Photosyn. Res. 31, 213.
Pechukas, P. (1994), Phys. Rev. Lett. 73, 1060.
Pelzer, K. M., G. B. Griffin, S. K. Gray, and G. S. Engel
(2012), J. Chem. Phys. 136, 164508.
Pflock, T. J., S. Oellerich, L. Krapf, J. Southhall, R. J.
Cogdell, G. M. Ullmann, and J. Kohler (2011), J. Phys.
Chem. B 115, 8821.
Pieper, J., M. Ratsep, I. Trostmann, H. Paulsen, G. Renger,
52
and A. Freiberg (2011), J. Phys. Chem. B 115, 4042.
Scheuring, S., R. P. Goncalves, V. Prima, and J. N. Sturgis
Plenio, M. B., J. Almeida, and S. F. Huelga (2013), J. Chem.
(2006), J. Mol. Biol. 358 (1), 83.
Phys. 139, 235102.
Scheuring, S., and J. N. Sturgis (2005), Science 309 (5733),
Plenio, M. B., and S. F. Huelga (2008), New J. Phys. 10,
484.
113019.
Scheuring, S., and J. N. Sturgis (2009), Photosyn. Res. 102,
Prandi, I. G., L. Viani, O. Andreussi,
and B. Mennucci
197.
(2016), J. Comput. Chem. 37, 981.
Schlau-Cohen, G. S., A. Ishizaki, and G. R. Fleming (2011a),
Prior, J., A. W. Chin, S. F. Huelga, and M. B. Plenio (2010),
Chem. Phys. 386, 1.
Phys. Rev. Lett. 105, 050404.
Prokhorenko, V. I., A. R. Holzwarth, F. R. Nowak, and T. J.
Aartsma (2002), J. Phys. Chem. B 106, 9923.
Schlau-Cohen, G. S., Q. Wang, J. Southall, R. J. Cogdell,
and W. E. Moerner (2011b), Proc. Natl. Acad. Sci., USA
110, 10899.
Pullerits, T., S. Hess, J. L. Herek, and V. Sundstrom (1997),
Schmidt am Busch, M., F. Muh, M. E.-A. Madjet,
and
J. Phys. Chem. B 101, 10560.
Purchase, R., and S. Volker (2009), Photosynth. Res. 101,
245.
Rancova, O., and D. Abramavicius (2014), J. Phys. Chem.
B 118, 7533.
T. Renger (2011), J. Phys. Chem. Lett. 2, 93.
Scholes, G. D. (2003), Annu. Rev. Phys. Chem. 54, 57.
Scholes, G. D. (2010), J. Phys. Chem. Lett. 1, 2.
Scholes, G. D., C. Curutchet, B. Mennucci, R. Cammi, and
J. Tomasi (2007), J. Phys. Chem. B 111, 6978.
Rancova, O., J. Sulskus,
and D. Abramavicius (2012), J.
Scholes, G. D., and G. R. Fleming (2000), J. Phys. Chem. B
Phys. Chem. B 116, 7803.
104, 1854.
Ratsep, M., R. E. Blankenship, and G. J. Small (1999), J.
Scholes, G. D., I. R. Gould, R. J. Cogdell, and G. R. Fleming
Phys. Chem. B 103, 5736.
(1999), J. Phys. Chem. B 103, 2543.
Ratsep, M., and A. Freiberg (2007), J. Lumin. 127, 251.
Ratsep, M., J. Pieper, K.-D. Irrgang, and A. Freiberg (2008),
Schroter, M., S. D. Ivanov, J. Schulze, S. P. Polyutov, Y. Yan,
T. Pullerits, and O. Kuhn (2015), Phys. Rep. 567, 1.
J. Phys. Chem. B 112, 110.
Read, E. L., G. S. Schlau-Cohen, G. S. Engel, J. Wen, R. E.
Blankenship, and G. R. Fleming (2008), Biophys. J. 95,
847.
Segatta, F., L. Cupellini, S. Jurinovich, S. Mukamel, M. Da-
por, S. Taioli, M. Garavelli, and B. Mennucci (2017), J.
Am. Chem. Soc. 139, 7558.
Sener, M., J. Strumpfer, A. Singharoy, C. N. Hunter, and
Rebentrost, P., R. Chakraborty,
and A. Aspuru-Guzik
K. Schulten (2016), eLife 5, e095421.
(2009a), J. Chem. Phys. 131, 184102.
Senn, H. M., and W. Thiel (2009), Angewandte Chemie (In-
Rebentrost, P., M. Mohseni, and A. Aspuru-Guzik (2009b),
ternational ed. in English) 48 (7), 1198.
Sharp, L. Z., D. Egorova, and W. Domcke (2010), J. Chem.
and
Phys. 132, 014501.
J. Phys. Chem. B 113, 9942.
Rebentrost, P., M. Mohseni,
I. Kassal, S. Lloyd,
A. Aspuru-Guzik (2009c), New J. Phys. 11, 033003.
Redfield, A. G. (1957), IBM Journal of Research and Devel-
opment 1, 19.
Renger, T. (2009), Photosyn. Res. 102, 471.
Renger, T., A. Klinger, F. Steinecker, M. Schmidt am Busch,
J. Numata, and F. Muh (2012), J. Phys. Chem. B 116,
14565.
Renger, T., and R. A. Marcus (2002), J. Chem. Phys. 116,
9997.
Renger, T., and V. May (1998), J. Phys. Chem. A 102, 4381.
Renger, T., V. May, and O. Kuhn (2001), Phys. Rep. 343,
137.
Renger, T., and F. Muh (2011), Photosyn. Res. 111 (1-2),
47.
Renger, T., and F. Muh (2013), Phys. Chem. Chem. Phys.
15, 3348.
Shi,, and Geva (2004), J. Chem. Phys. 120, 10647.
Shim, S., P. Rebentrost, S. Valleau, and A. Aspuru-Guzik
(2012), Biophys. J 102, 649.
Silbey, R. (1976), Annu. Rev. Phys. Chem. 27, 203.
Sisto, A., C. Stross, M. W. van der Kamp, M. O. S. M.-
S. G. T. Johnson, E. G. Hohenstein, F. R. Manby, D. R.
Glowacki, and T. J. Martinez (2017), Phys. Chem. Chem.
Phys. 19, 14924.
Skochdopole, N., and D. A. Mazziotti (2011), J. Phys. Chem.
Lett. 2, 2989.
Steindal, A. H., K. Ruud, L. Frediani, K. Aidas, and J. Kong-
sted (2011), The Journal of Physical Chemistry B 115 (12),
3027.
Stross, C., M. W. van der Kamp, T. A. A. Oliver, J. N. Har-
vey, N. Linden, and F. R. Manby (2016), J. Phys. Chem.
B 120, 11449.
Ritschel, G., J. Roden, W. T. Strunz, and A. Eisfeld (2011),
Strumpfer, J., and K. Schulten (2009), J. Chem. Phys. 131,
New. J. Phys. 13, 113034.
225101.
Ritz, T., S. Park, and K. Schulten (2001), J. Phys. Chem. B
Strumpfer, J., M. Sener, and K. Schulten (2012), J. Phys.
105, 8259.
Chem. Lett. 3, 536.
Rosenbach, P., J. Cerrillo, S. F. Huelga, J. Cao, and M. B.
Sturgis, J., J. D. Tucker, J. D. Olsen, C. N. Hunter, and
Plenio (2016), New J. Phys. 18, 023035.
Saga, Y., Y. Shibata, and H. Tamiaki (2010), J. Photochem.
Photobiol. C: Photochem. Rev. 11, 15.
Sarovar, M., A. Ishizaki, G. R. Fleming, and K. B. Whaley
(2010), Nat. Phys. 6, 462.
R. A. Niederman (2009), Biochemistry 48, 3679.
Sumi, H. (1999), J. Phys. Chem. B 103, 252.
Sumino, A., T. Dewa, T. Noji, Y. Nakano, N. Watanabe,
R. Hildner, N. Bosch, J. Kohler, and M. Nango (2013), J.
Phys. Chem. B 117, 10395.
Savikhin, S., D. R. Buck, and W. S. Struve (1997), Chem.
Sundstrom, V., T. Pullerits, and R. van Grondelle (1999), J.
Phys. 223, 303.
Savikhin, S.,
and W. S. Struve (1994), Biochemistry 33,
11200.
Savikhin, S., and W. S. Struve (1996), Photosyn. Res. 48,
271.
Phys. Chem. B 103, 2327.
Tanimura, Y. (2006), J. Phys. Soc. Jpn. 75, 082001.
Tanimura, Y. (2015), J. Chem. Phys. 142, 144110.
Tanimura, Y., and S. Mukamel (1993), Phys. Rev. E 47, 118.
Tempelaar, R., T. L. C. Jansen, and J. Knoester (2014), J.
53
Phys. Chem. B 118, 12865.
Phys. Chem. C 118 (43), 24854.
Thyrhaug, E., K. Z´ıdek, J. Dost´al, D. B´ına, and D. Zigmantas
Wang, X., G. Titschel, S. Wuster, and A. Eisfeld (2015),
(2016), J. Phys. Chem. Lett. 7, 1653.
Phys. Chem. Chem. Phys. 17, 25629.
Timpmann, K., M. Ratsep, C. N. Hunter, and A. Freiberg
(2000a), J. Phys. Chem. B 104, 10581.
Timpmann, K., N. W. Woodbury, and A. Freiberg (2000b),
J. Phys. Chem. B 104, 9769.
Tiwari, V., W. K. Peters, and D. M. Jonas (2013), Proc.
Natl. Acad. Sci. USA 110, 1203.
Wendling, M., M. A. Przyjalgowski, D. Gulen, S. I. E. Vulto,
T. J. Aartsma, R. van Grondelle, and H. van Amerongen
(2002), Photosyn. Res. 71, 99.
Wilk, K. E., S. J. Harrop, L. Jankova, D. Edler, G. Keenan,
F. Sharples, R. Hiller, and P. M. G. Curmi (1999), Proc.
Natl. Acad. Sci., USA 96, 8901.
Tomasi, J., B. Mennucci, and R. Cammi (2005), Chem. Rev.
Wilkins, D.,
and N. S. Dattani (2015), J. Chem. Theory
105, 2999.
Trinkunas, G., J. L. Herek, T. Polivka, V. Sundstrom, and
T. Pullerits (2001), Phys. Rev. Lett. 86, 4167.
Trinkunas, G., O. Zerlauskiene, V. Urboniene, J. Chmeliov,
and L. Valkunas (2012), J. Phys.
A. Gall, B. Robert,
Chem. B 116, 5192.
Tronrud, D. E., J. Wen, L. Gay,
and R. E. Blankenship
(2009), Photosynth. Res. 100, 79.
Tsuneda, T., and K. Hirao (2014), Wiley Interdisciplinary
Reviews: Computational Molecular Science 4 (4), 375.
Tully, J. C. (2012), J. Chem. Phys. 137, 22A301.
Turner, D. B., R. Dinshaw, K. K. Lee, M. S. Belsley, K. E.
Wilk, P. M. G. Curmi, and G. D. Scholes (2012), Phys.
Chem. Chem. Phys. 14, 4857.
Urboniene, V., O. Vrublevskaja, G. Trinkunas, A. Gall,
B. Robert, and L. Valkunas (2007), Biophys. J. 93, 2188.
and A. Asupuru-Guzik (2012), J.
Valleau, S., A. Eisfeld,
Chem. Phys. 137, 224103.
van der Vegte, C. P., J. D. Prajapati, U. Kleinekathoefer,
J. Knoester, and T. L. C. Jansen (2015), J. Phys. Chem.
B 119 (4), 1302.
Viani, L., M. Corbella, C. Curutchet, E. J. O'Reilly, A. Olaya-
Castro, and B. Mennucci (2014), Phys. Chem. Chem. Phys.
16, 16302.
Viani, L., C. Curutchet, and B. Mennucci (2013), J. Phys.
Chem. Lett. 4, 372.
Vlaming, S. M., and R. J. Silbey (2012), J. Chem. Phys. 136,
055102.
Voityuk, A. A. (2013), J Phys Chem C 117 (6), 2670.
Voityuk, A. A. (2014), J. Chem. Phys. 140 (24), 244117.
Vulto, S. I. E., M. A. de Baat, R. J. W. Louwe, H. P. Per-
mentier, T. Neef, M. Miller, H. van Amerongen, and T. J.
Aartsma (1998), J. Phys. Chem. B 102, 9577.
Vulto, S. I. E., M. A. de Baat, S. Neerken, F. R. Nowak,
H. van Amerongen, J. Amesz, and T. J. Aartsma (1999),
J. Phys. Chem. B 103, 8153.
Vura-Weis, J., M. D. Newton, M. R. Wasielewski, and J. E.
Subotnik (2010), J. Phys. Chem. C 114, 20449.
Wan, Y., A. Stradomska, S. Fong, Z. Guo, R. D. Schaller,
G. P. Wiederrecht, J. Knoester, and L. Huang (2014), J.
Comput. 11, 3411.
Wong, C. Y., R. M. Alvey, D. B. Turner, K. E. Wilk, D. A.
Bryant, P. M. G. Curmi, R. J. Silbey, and G. D. Scholes
(2012), Nat. Chem. 4, 396.
Wu, H.-M., M. Ratsep, R. Jankowiak, R. J. Cogdell, and
G. J. Small (1997), J. Phys. Chem. B 101, 7641.
Wu, H.-M., and G. J. Small (1998), J. Phys. Chem. B 102,
888.
Wu, J. L., F. Liu, J. Ma, R. J. Silbey, and J. Cao (2012), J.
Chem. Phys. 137, 174111.
Xiong, S.-J., Y. Xiong, and Y. Zhao (2012), J. Chem. Phys.
137, 094107.
Yang, C.-H., and C.-P. Hsu (2013), J. Chem. Phys. 139 (15),
154104.
Yang, G., N. Wu, T. Chen, K. Sun, and Y. Zhao (2012), J.
Phys. Chem. C 116, 3747.
Yang, M., R. Agarwal, and G. R. Fleming (2001), J. Pho-
tochem. Photobiol. A: Chemistry 142, 107.
Yang, M., and G. R. Fleming (2002), Chem. Phys. 275, 355.
Yeh, S.-H., J. Zhu, and S. Kais (2012), J. Chem. Phys. 137,
084110.
You, Z.-Q.,
and C.-P. Hsu (2014), Int J Quantum Chem
114 (2), 102.
Zerlauskiene, O., G. Trinkunas, A. Gall, B. Robert, V. Ur-
boniene, and L. Valkunas (2008), J. Phys. Chem. B 112,
15883.
Zhang, M.-L., B. J. Ka, and E. Geva (2006), J. Chem. Phys.
125, 044106.
Zhang, W. M., T. Meier, V. Chernyak,
and S. Mukamel
(1998), J. Chem. Phys. 108, 7763.
Zheng, F., S. Fernandez-Alberti, S. Tretiak, and Y. Zhao
(2017), J. Phys. Chem. 121, 5331.
Zheng, X., R. X. Xu, J. Xu, J. S. Jin, J. Hu, and Y. J. Yan
(2012), Prog. Chem. 24, 1129.
Zigmantas, D., E. L. Read, T. Mancal, T. Brixner, A. T.
Gardiner, R. J. Cogdell, and G. R. Fleming (2006), Proc.
Natl. Acad. Sci., USA 103, 12672.
Zwier, M. C., J. M. Shorb, and B. P. Krueger (2007), J.
Comput. Chem. 28, 1572.
|
1908.08376 | 1 | 1908 | 2019-08-22T13:44:36 | Encounter rates between bacteria and small sinking particles | [
"physics.bio-ph",
"physics.flu-dyn"
] | Bacteria in aquatic environments often interact with particulate matter. A key example is bacterial degradation of marine snow responsible for carbon export from the upper ocean in the biological pump. The ecological interaction between bacteria and sinking particles is regulated by their encounter rate, which is therefore important to predict accurately in models of bacteria-particle interactions. Models available to date cover the diffusive encounter regime, valid for sinking particles larger than the typical run length of a bacterium. The majority of sinking particles, however, are small, and the encounter process is then ballistic rather than diffusive. In the ballistic regime, the shear generated by the particle's motion can be important in reorienting bacteria and thus determining the encounter rate, yet the effect of shear is not captured in current encounter rate models. Here, we combine analytical and numerical calculations to quantify the encounter rate between sinking particles and non-motile or motile microorganisms in the ballistic regime, explicitly accounting for the hydrodynamic shear created by the particle and its coupling with microorganism shape. We complement results with selected experiments on non-motile diatoms. We find that the shape-shear coupling has a considerable effect on the encounter rate and encounter location through the mechanisms of hydrodynamic focusing and screening, whereby elongated microorganisms preferentially orient normally to the particle surface downstream of the particle (focusing) and tangentially to the particle surface upstream of the particle (screening). We study these mechanisms as a function of the key dimensionless parameters: the ratio of particle sinking speed to microorganism swimming speed, the ratio of particle radius to microorganism length, and the microorganism's aspect ratio. | physics.bio-ph | physics |
Encounter rates between bacteria and small sinking particles
Jonasz S(cid:32)lomka,1 Uria Alcolombri,1 Eleonora Secchi,1 Roman Stocker,1 and Vicente I. Fernandez1
1Institute of Environmental Engineering, Department of Civil,
Environmental and Geomatic Engineering, ETH Zurich
(Dated: August 23, 2019)
Bacteria in aquatic environments often interact with particulate matter. A key example is bacte-
rial degradation of marine snow responsible for carbon export from the upper ocean in the biolog-
ical pump. The ecological interaction between bacteria and sinking particles is regulated by their
encounter rate, which is therefore important to predict accurately in models of bacteria-particle in-
teractions. Models available to date cover the diffusive encounter regime, valid for sinking particles
larger than the typical run length of a bacterium. The majority of sinking particles, however, are
small, and the encounter process is then ballistic rather than diffusive. In the ballistic regime, the
shear generated by the particle's motion can be important in reorienting bacteria and thus determin-
ing the encounter rate, yet the effect of shear is not captured in current encounter rate models. Here,
we combine analytical and numerical calculations to quantify the encounter rate between sinking
particles and non-motile or motile microorganisms in the ballistic regime, explicitly accounting for
the hydrodynamic shear created by the particle and its coupling with microorganism shape. We
complement results with selected experiments on non-motile diatoms. We find that the shape -- shear
coupling has a considerable effect on the encounter rate and encounter location through the mech-
anisms of hydrodynamic focusing and screening, whereby elongated microorganisms preferentially
orient normally to the particle surface downstream of the particle (focusing) and tangentially to the
particle surface upstream of the particle (screening). We study these mechanisms as a function of
the key dimensionless parameters: the ratio of particle sinking speed to microorganism swimming
speed, the ratio of particle radius to microorganism length, and the microorganism's aspect ratio.
We find that non-motile elongated microorganisms are screened from sinking particles in ballistic
interactions because shear aligns them tangentially to the particle surface. As a result, the en-
counter rate is reduced by a factor proportional to the square of the microorganism aspect ratio as
compared to a spherical microorganism. For motile elongated microorganisms, hydrodynamic fo-
cusing increases the encounter rate approximately twofold compared to the case without shear when
particle sinking speed is similar to microorganism swimming speed, whereas for very quickly sinking
particles hydrodynamic screening can reduce the encounter rate below that of non-motile microor-
ganisms. We apply these results to predict the encounter rates of submillimetric marine particles
with bacteria under natural ocean conditions. In this size range, which covers the most abundant
marine particles, the shear-induced reorientation competes with randomization of the swimming di-
rection due to Brownian effects and run-and-tumble motility. We present comprehensive maps that
connect the ballistic and diffusive limits and yield the encounter rate as a function of shape, motility
and particle characteristics. Overall, our results indicate that hydrodynamic shear is important to
consider in the colonization and ultimately degradation of marine particles, with a direct impact
on which particles are colonized and where. Shear should thus be taken into account to predict the
dynamics of settling particles responsible for the large carbon flux in the ocean's biological pump.
I.
INTRODUCTION
Encounters involving small particles suspended in a
fluid underpin many industrial, physical and biological
processes. In papermaking, too high a collision rate be-
tween cellulose fibers leads to excessive fiber flocculation
and poor paper quality [1].
In the atmosphere, pre-
cipitation formation relies on encounters between water
droplets in clouds under the combined action of grav-
ity and turbulence [2].
In the ocean, encounter rates
between microscopic phytoplankton following a phyto-
plankton bloom determine the formation of marine snow
responsible for the biological pump, the vertical flux of
carbon from the upper ocean to its depths [3]. Liv-
ing organisms extend the complexity of the encounter
processes occurring in non-motile systems by additional
mechanisms. Microorganisms and plankton dwelling in
the oceans can navigate through water in search of food
and motility greatly enhances the encounter rates of these
microscopic organisms with resource patches [4]. Com-
pared to non-motile microorganisms, whose encounter
rate is proportional to the low diffusivity associated with
Brownian motion, motile microorganisms have a much
higher (often, 100- to 1000-fold) encounter rate, since
their motility effectively enhances diffusivity [4]. Of par-
ticular importance for the biogeochemical cycles of car-
bon in the ocean are the encounters between bacteria and
sinking particles of organic matter. Once attached to a
particle, bacteria can grow on it and solubilize it, thus re-
ducing the flux of carbon to the deep ocean [5, 6], a fun-
damental process in climate-relevant carbon dynamics.
Accurate models of the encounter rate between bacteria
and particles valid across a wide range of particle sizes
are thus important to estimate the role of bacteria in the
carbon pump. To date, however, encounter rate formu-
lations have focused on the diffusive regime suitable for
large particles. Here, we study the encounter between mi-
croorganisms and sinking particles in the ballistic limit,
relevant for the most abundant small particles, with fo-
cus on the impact of fluid flow and the associated shear
generated by the particle on the encounter rates.
Theoretical estimates of the encounter rates between
microorganisms and sinking particles have thus far pri-
marily built on modeling microorganisms as spherical col-
loids and motility as a diffusive process [7, 8]. These sim-
plifying assumptions map the microbial encounter with
particles onto the classical problem of heat and mass
transfer [8, 9]. By construction, this approach assumes
particles are larger than the run length of a bacterium.
Since the latter is of the order of tens to hundred of mi-
crons [8, 10], the diffusive approximation is limited to
particles larger than several hundred microns. Yet, due
to the power-law nature of the marine particle size spec-
trum, the most abundant particles in the ocean have sizes
below hundred microns [11]. In this increasingly ballis-
tic regime, the coupling between the flow generated by
the particle and the swimming of bacteria may domi-
nate the bacterial orientational dynamics, in contrast to
the diffusive regime [12]. There is substantial experimen-
tal and theoretical evidence that fluid velocity gradients
(shear) can dramatically modify the swimming trajecto-
ries of microorganisms [13 -- 15]. A primary mechanism
is shear-induced reorientation, whereby the torque as-
sociated with fluid velocity gradients reorients microor-
ganisms and thus impacts their swimming direction and
where they end up in the flow. For example, a sim-
ple parabolic flow can lead to shear-trapping and bac-
terial accumulation near microchannel walls [14]. Shear-
induced reorientation is a general phenomenon, applica-
ble to any elongated bacteria that swim in flow, yet its
impacts on the fundamental problem of the encounter
rate between microorganisms and sinking aggregates in
the ballistic range has to date not been considered.
Here, we combine analytical and numerical calcula-
tions with experiments to study encounters between non-
motile and motile microorganisms and sinking particles
in the ballistic regime, with focus on how the flow cre-
ated by the particle affects bacterial trajectories and ul-
timately the encounter rates. For the classical Stokes
flow around a sphere, we show analytically that the ori-
entational dynamics of elongated bacteria - unlike spher-
ical particles - break the fore -- aft symmetry of the flow
streamlines, with major consequences on encounter rates
and attachment location. Non-motile elongated bacteria
orient tangentially to the particle surface as they pass
by the particle, which reduces their encounter rate by
a factor proportional to the square of the bacterial as-
pect ratio. For motile elongated bacteria, the encounter
rate is very sensitive to the particle sinking speed rela-
tive to the bacterial swimming speed. When both speeds
are comparable, shear increases the encounter rate about
twofold and leads to preferential attachment to the lee-
ward side of the particle. For rapidly sinking particles,
shear screens motile bacteria from the sinking particle
2
FIG. 1: The ballistic model of the encounter between bacteria
and sinking particles includes the impact of shear on bacterial
trajectories. (a) Spherical particle sinks under gravity with
speed given by the Stokes law (1) and induces the Stokes
flow (2) around it. Bacteria are modeled as self-propelled el-
lipsoids of aspect ratio α with the center of mass x(t) and
the tail-to-head orientation p(t) obeying Eq. (3). (b) Rep-
resentative trajectories for a bacterium starting at [x, y, z] =
R[2,−2,−2] with a random initial orientation (U/Ub = 3 and
α = 10). Red (black) trajectories correspond to interceptions
(misses). Interceptions are characterized by the landing posi-
tion on the particle as well as the initial orientation. All such
initial orientations define the interception probability starting
at the given position [red points in the inset, Eq. (4)].
and surprisingly, the encounter rate drops far below the
limit corresponding to non-motile bacteria.
This work is organized as follows: we introduce the
model of the encounter process and define the relevant
observables in Section II. To quantify the impact of shear
on bacterial orientation, we classify the asymptotic con-
figurations that ellipsoids assume in general flows and
then apply the results to the Stokes flow around a sphere
in Section III. The encounter of non-motile and motile
microorganisms with sinking particles is studied in Sec-
tions IV and V. We discuss the biophysical consequences
of our mechanistic description of the encounter process
in Section VI and draw conclusions in Section VII.
II. MODEL
We model a marine snow particle as a sphere sinking
in a quiescent fluid and bacteria as elongated and self-
propelled ellipsoids (Section II A). The encounter process
is quantified through encounter rate, encounter efficiency
and distribution of interception locations (Section II B).
A. Equations of motion
The most abundant marine snow particles in the ocean
have sizes in the range up to several hundred microns [11]
and sinking speeds up to about a millimeter per sec-
ond [16], which gives Reynolds number up to about 0.1.
x/R02-2y/R2-20z/R2-20Umisshit (b)1.95-2.05-2-2-2.05-1.952.051.952hitinitial orientations:particle shadow2011-1001-1x(t)p(t)(a)2-6-40-2ρ/RRθUz/R2lbUb(cid:17)θ,(2)
(cid:16)
(cid:17)
(cid:16)
In this viscosity-dominated regime, the gravitational and
viscous forces on the particle balance, implying that a
spherical particle of radius R sinks at the constant ter-
minal speed given by the Stokes law
U =
2
9
ρp/ρw − 1
ν
gR2,
(1)
where ρp and ρw are the densities of the particle and
water, respectively, ν is the kinematic viscosity of water
and g is the gravitational acceleration. In the reference
frame fixed at particle and moving with it [Fig. 1(a)], the
flow is described by the classic Stokes flow
v = U cos θ
1 +
R3
2r3 − 3R
2r
r − U sin θ
1 − R3
4r3 − 3R
4r
where U is the sinking speed given by Eq. (1).
We model a bacterium as a small self-propelled elon-
gated ellipsoid characterized by three parameters: length
lb, aspect ratio α and swimming speed Ub. The position
and orientation of the bacterium at time t are given by
x(t) and p(t), where the latter (unit) vector points from
the bacterial tail to its head [Fig. 1(b)]. The dynamics
of x and p are governed by
x = Ubp + v,
p = (I − ppT)(γE + W )p.
(3a)
(3b)
Eq. (3a) states that the total bacterial velocity x is a su-
perposition of self-propulsion with speed Ub in the direc-
tion p and the flow v (2) around the particle. Eq. (3b)
is the classic Jeffrey equation for the orientational dy-
namics of ellipsoids in flow [17]. The tensors E and W
are the symmetric and anti-symmetric parts of the ve-
locity gradient Aij = ∂jvi. The bacterial aspect ratio
enters the dynamics (3) through the shape parameter
γ = (α2 − 1)/(α2 + 1); it vanishes for spheres, is positive
for elongated organisms and negative for oblate ones.
B. Physical observables
Let p(x, p) be the probability of an encounter between
the sinking particle and a bacterium starting at the initial
position x with head pointing in the direction p. For the
ballistic model (3), p is either zero or one since the initial
condition (x, p) determines a unique bacterial trajectory;
when Eq. (3) is supplemented with rotational diffusion,
p can take a range of values between 0 and 1. Averaging
over random orientations yields the encounter probability
P (x) for an initial position x
(cid:90)
P (x) =
dpp(x, p).
(4)
Intuitively, P (x) is the relative solid angle extended by
initial bacterial orientations that lead to the intercep-
tion (red area in the inset of Fig. 1(b)]. Let (z, ρ, φ) be
the cylindrical coordinate system with origin fixed at the
3
(cid:90) ∞
sinking particle. Due to rotational symmetry around the
z-axis, we have P (x) = P (z, ρ). To define the encounter
rate and interception efficiency, suppose that the sink-
ing particle enters a region of uniform concentration n
of randomly oriented bacteria. Let N (t, z) be the total
number of encounters with bacteria that at time t are lo-
cated at a z-plane below the sinking particle (upstream of
the particle, z < 0) and collide with the particle at some
later time. In a short interval (t, t + dt), the change in N
due to the encounters with bacteria with initial positions
0 P (z, ρ)ρdρ.
Therefore, for a constant sinking speed, the encounter
rate d N /dt(z) is independent of time and is given by
in the thin sheet (z, z + U dt) is 2πnU dt(cid:82) ∞
d N /dt(z) = 2πnU
(5)
For a z-plane far away from the sinking particle z (cid:29) R
the fluid is practically undisturbed, making it meaningful
to define the z-independent encounter rate N = dN/dt
P (z, ρ)ρdρ.
0
N = d N /dt(z → ∞).
(6)
In simulations, we fix the starting plane at z = −6R,
which amounts to making the approximation N ≈
d N /dt(z = −6R). To scale out the concentration n, we
often focus on N /n, the 'encounter rate kernel' [8].
To further scale out factors intrinsic to the sinking par-
ticle, the radius R and velocity U , we follow the notation
used in filtration literature and define the dimensionless
interception efficiency η as the ratio of volume cleared
and volume swept by the particle [9]
η =
N /n
πR2U
=
2
R2
P (ρ)ρdρ.
(7)
Intuitively, η = 1 means that the sinking particle collects
bacteria from a volume of water equal to the volume of
the cylinder the particle sweeps. For small non-motile
colloids, we expect η (cid:28) 1 because the colloids are con-
strained to the flow streamlines, which limits the inter-
ception to a narrow region near the particle centerline,
the 'stagnation line' [Fig. 4(a) and Section IV A].
In addition to computing the encounter rate and en-
counter efficiency, we will quantify the location on the
sinking particle where the bacteria land. Let ξ(θ, φ) be
the distribution of the interception locations, where θ
and φ are the colatitude and the azimuth coordinates
on the particle, respectively. We normalize ξ(θ, φ) as
the probability density function over the unit sphere,
ξ(θ, φ)dΩ = 1, where dΩ = sin θdθdφ. Rotational
0
symmetry implies that ξ(θ, φ) = ξ(θ). Finally, the mean
interception colatitude is
(cid:82) 2π
(cid:82) π
0
(cid:90) ∞
0
(cid:90) π
(cid:90) 2π
(cid:104)θ(cid:105) =
(cid:90) π
0
0
θξ(θ) sin θdθ.
θξ(θ, φ)dΩ = 2π
(8)
For example, 0° < (cid:104)θ(cid:105) < 90° (northern hemisphere, down-
stream) implies preferential leeward attachment, while
90° < (cid:104)θ(cid:105) < 180° (southern hemisphere, upstream) indi-
cates attachment to the front.
0
4
time limit of the Jeffrey Eq. (3b). Previous studies fo-
cused on special cases with A derived from, for exam-
ple, simple shear or rotational flows; in the former case
the dynamics collapse onto one of the many degenerate
limit cycles, the well-known Jeffrey orbits [17, 18]. For a
random A, neglecting marginal cases, two scenarios are
possible: either a rod asymptotically points towards the
direction of the largest effective deformation rate or it
rotates in a certain plane. The first possibility has been
known [18] and corresponds to the rate of strain E out-
competing the rate of rotation W . The second scenario
complements the study in [19] and generalizes the Jef-
frey orbits to generic rotational flows and arises when W
dominates over E. Detailed derivations are given in the
Appendices A 1 and A 2.
The velocity gradient tensor A has nine components,
eight of which are independent for an incompressible
flow (since Aii = ∂ivi = 0). The symmetric part of
A, the rate of strain E, describes the rate at which the
fluid stretches and compresses [20]. The antisymmetric
part W represents the fluid rate of rotation and is deter-
mined by the vorticity ω = ∇ × u as Wij = − 1
2 ijkωk.
Given A, it is the weighted sum Aγ = γE + W that en-
ters the Jeffrey Eq. (3b), where γ is the shape parameter
γ = (α2− 1)/(α2 + 1) determined by the organism aspect
ratio α. In this notation, the Jeffrey Eq. (9) reads
p = (I − ppT)Aγp.
(9)
Eq. (9) is a dynamical system on the unit sphere of ori-
entations (Fig. 2). We first discuss the case of spherical
microorganisms (α = 1, γ = 0) and then describe in de-
tail the response of elongated bacteria (α > 1, γ > 0);
the case of oblate microorganisms (α < 1, γ < 0) is dual
to that of elongated ones.
For spherical microorganisms, the shape parameter
vanishes (γ = 0) and Eq. (9) simplifies to
p =
ω × p,
1
2
(10)
where ω is the vorticity. Thus, spherical microorganisms
respond to the fluid rotation but are unaffected by the
fluid straining motion. Eq. (10) can be solved exactly in
this case [18]: for a given initial orientation p(0), the so-
lutions of Eq. (10) correspond to p(t) rotating around the
vorticity vector ω and in the same sense as ω [Fig. 2(a)].
The angle between p(t) and ω is fixed by the initial orien-
tation and the rotation rate is (cid:107)ω(cid:107)/2. Equivalently, the
solutions to Eq. (10) on the unit sphere of orientations
with ω pointing along the z-axis are given by the circles
of a fixed latitude.
Elongated or oblate microorganisms respond to both,
the fluid rate of strain and rate of rotation.
In this
case, it appears impossible to find analytical solutions
to Eq. (9); instead, standard dynamical system theory
helps to identify the long-time response. Since the fixed
points of Eq. (9) are given by the real eigenvectors of
Aγ [18], it is the eigendecomposition of Aγ that deter-
mines the asymptotic response. For a random Aγ, two
FIG. 2:
In flow, spherical bacteria rotate around the vorticity
vector (a), whereas elongated or oblate ones eventually point
in the direction of the largest effective deformation rate (b)
or rotate in a certain plane (c). (a-c) Phase portraits of the
Jeffrey Eq. (9) for the bacterial tail-head vector p(t) reorient-
ing under the velocity gradient A for different A and aspect
ratios α. (a) Spherical bacteria always rotate around the vor-
ticity ω (z-axis). (b,c) Depending on the flow being strain- or
rotation-dominated, elongated bacteria (or flat disks) eventu-
ally point along the direction of the largest effective deforma-
tion rate [z-axis in (b)] or rotate in the plane perpendicular
to the real eigenvector of (Aγ)T [equator in (c)]. (d) Time se-
ries of the components of p(t) for the case shown in (c). The
rotation in the x − y plane is nonuniform: the rod acceler-
ates when approaching the straining direction but slows down
when near the axis of compression; the rotation frequency
is given by the imaginary part of the complex eigenvalue
of Aγ. Parameters: A = [0,−1/2, 0; 1/2, 0, 0; 0, 0, 0], α =
1 (a), A = [−3/4, 0, 0; 0,−1/4, 0; 0, 0, 1], α = ∞ (b) and
A = [1/2, 1, 0;−1/4, 1/2, 0; 0, 0,−1], α = ∞ (c).
III. ELLIPSOIDS IN FLOW
Shear preferentially reorients rods and disks, such as
elongated bacteria or flat diatoms, along certain direc-
tions. Depending on the flow being strain- or rotation-
dominated, ellipsoids eventually point in the direction
of the largest deformation rate or rotate in a certain
plane (Section III A). Applying this classification to the
Stokes flow induced by the sinking particle reveals that
ellipsoids - unlike spheres - break the fore-aft symmetry
of the flow streamlines (Section III B). This quasi-static
picture will be essential to rationalize the subsequent sim-
ulations of the encounter problem as it underpins the
phenomena of hydrodynamic focusing and screening.
A. Ellipsoids in general flows
The asymptotic orientation of a nonspherical microor-
ganism held fixed in flow but free to reorient under the
action of the velocity gradient A follows from the long-
(a)(b)1-10pz1-10py01-1px10Reorientation speed ║p║.01-1px1-10py1-10pz(c)(d)pi-110pxpypz013Time2(d)1-10pz1-10py01-1pxcases are possible: either Aγ has three real eigenvalues
or one real eigenvalue and two complex conjugate eigen-
values. In the first case, the eigenvector corresponding
to the largest positive eigenvalue is an attractive fixed
point of (9) [Fig. 2(b)]. In the second case, when the real
eigenvalue is positive, the corresponding eigendirection is
still attractive, but once this eigenvalue is negative, the
eigendirection becomes unstable and the dynamics col-
lapse onto a limit cycle [Fig. 2(c)]. We next discuss these
asymptotic scenarios in more detail.
= 3
Let λi and λi, where i = 1, 2, 3, be the eigenvalues and
eigenvectors of Aγ. The incompressibility of the flow re-
quires that λ1 + λ2 + λ3 = 0. When all λi's are real, the
Jeffrey Eq. (9) has three pairs of fixed points correspond-
ing to p = ±λi. Assuming that λ1 < λ2 < λ3, the fixed
points are respectively: a repulsive node (λ1), a saddle
(λ2), and an attractive node (λ3). As a consequence, a
random initial orientation eventually collapses onto the
stable direction p = ±λ3 [Fig. 2(b)]. The timescale τλ3
associated with this reorientation is estimated as inverse
of the average of the eigenvalues of the linearized ver-
sion of Eq. (9) near the fixed point λ3 and is given by
2 λ3. Since the system orients along ±λ3, the case
τ−1
λ3
when all λi's are real corresponds to the rate of strain E
outcompeting the rate of rotation W .
When Aγ has a pair of complex eigenvalues λ1 and λ∗
1,
and a real eigenvalue λ3, the eigenvectors corresponding
1} are complex, implying that there are only two
to {λ1, λ∗
fixed points given by p = ±λ3. When λ3 > 0 (stretch-
ing), the fixed point λ3 is an attractive spiral and repre-
sents the asymptotic direction. The appearance of com-
plex eigenvalues signals the raising importance of the rate
of rotation W , but when λ3 > 0, straining still dominates
the response. However, when λ3 < 0 (compression), the
fixed point λ3 becomes a repulsive spiral and a stable
limit cycle emerges. The timescale τλ3 of spiraling onto
2λ3. The limit cycle cor-
or away from λ3 is τ−1
responds to a great circle; the circle lies in the plane
(cid:48)
3, the eigenvector of the
with normal direction given by λ
transpose matrix Aγ T with eigenvalue λ3. Thus, when
λ3 < 0, the asymptotic state of Eq.(9) corresponds to p
(cid:48)
3 [Fig. 2(c)]. The an-
rotating in the plane normal to λ
gular frequency of the rotation is given by the imaginary
part of the complex eigenvalue λ1 [Fig. 2(d)].
= 3
λ3
The above analysis applies to a velocity gradient Aγ
under the assumption that all its eigenvalues are differ-
ent; a separate analysis is required in the degenerate case.
We next study A derived from the Stokes flow around a
sinking sphere.
B. Ellipsoids in the Stokes flow
The above classification of the orientational response
of a microorganism is now specified to the velocity gradi-
ent derived from the Stokes flow (2). Physically, we de-
scribe a bacterium rotating freely under shear but with
the center of mass fixed at some position. Self-propulsion
5
and advection are still not included and this simplifi-
cation makes analytical progress possible. The long-
time orientation depends on bacterial position: spherical
bacteria respond identically upstream and downstream
of the sinking particle [Fig. 3(a)], whereas nonspherical
ones break the streamline fore -- aft symmetry [Fig. 3(b,c)].
This symmetry breaking leads to hydrodynamic focus-
ing and screening, crucial shear-induced mechanisms that
impact the full encounter problem.
Spherical bacteria respond solely to the fluid vorticity
(Section III A), which for the Stokes flow (2) reads
ω = − 3
2
U R
sin θ
r2
φ.
(11)
It follows from Eq. (11) that bacteria rotate around the
azimuth φ, [Fig. 3(a)]. The rotation rate decays with the
square of the distance from the particle; it is strongest
to the side of the particle, near the equator θ = π/2 and
vanishes near the stagnation lines θ = 0 and θ = π. In
particular, the response of spherical bacteria preserves
the fore-aft symmetry of the flow streamlines: at a fixed
distance r, the bacterial rotation is identical at colati-
tudes θ and π − θ. As discussed next, this symmetry is
broken for elongated or oblate microorganisms.
We focus on perfect rods of infinite aspect ratio α →
∞, for which the shape factor γ = 1 in the Jeffrey
Eq. (3b) [Fig. 3(b)]. Note that moderate elongation gives
γ close to unity: for an aspect ratio α = 10, γ ≈ 0.98.
The analysis of the response of disks is dual to that of
rods and we only state the results [Fig. 3(c)]. For per-
fect rods, the rates of strain and rotation, E and W , are
weighted equally in the Jeffrey Eq. (9), which reduces to
(Aγ=1 = A)
p = (I − ppT)Ap,
(12)
where A is the velocity gradient derived from the Stokes
flow (2). For brevity, we take U = 1 and R = 1. In spher-
ical coordinates {r, θ, φ} and in the usual basis of unit
vectors {r, θ, φ}, the entries of A read (Appendix A 3,
see also [21])
Aij = F (r, θ)
2
tan θ
0
−β(r) tan θ −1
0
0 −1
0
(13)
where F = 3 cos θ(r−2−r−4)/4 and β = (r2 +1)/(r2−1).
To classify the response of rods in the manner outlined
in Section III A, we find the eigenvalues of A
λ1,2 =
F
2
9 − 4β tan2 θ
λ3 = −F,
,
(14)
(cid:17)
,
(cid:16)
1 ±(cid:113)
−3 ±(cid:112)
and the corresponding eigenvectors
λ1,2 = [1,
9 + 4β tan2 θ
2 tan θ
, 0], λ3 = φ. (15)
From the sing change under the square root in Eq. (14),
it follows that the regions in the fluid in which A has
6
FIG. 3: Nonspherical microorganisms break the fore-aft symmetry of the Stokes flow around a sinking particle as revealed
by their asymptotic orientation when held fixed in the flow. The velocity gradient A [Eq. (13)] determines the local long-
time orientation of the bacterial tail-to-head vector p(t) (Fig. 2). (a) Spherical bacteria respond symmetrically upstream and
downstream of the particle: p rotates around the vorticity ω ∝ φ and in the same sense as ω (arrow); the broken line indicates
that there are many possible orbits [Fig. 2(a)]. (b) Perfect rods exhibit three regions of different asymptotic orientations.
In regions I and III the rate of strain outcompetes the vorticity [Fig. 2(b)], whereas the vorticity dominates in the region
II [Fig. 2(c)]. Specifically, in the upstream region I, bacteria eventually point along the azimuth p = ± φ. The two subregions
inside the region I differ only by how the asymptotic orientation is attained (attractive node vs spiral). In region II, p eventually
rotates in the plane perpendicular to φ [Fig. 2(c)]. The color code shows the rotation rate normalized by (cid:107)ω(cid:107)/2; the sense of
rotation is the same as ω (solid line and arrow). In the downstream region III, p orients along the director field (white lines)
defined by λ1 in Eq. (15). (c) For perfect disks, the response is a reflection of the case of perfect rods. (d) Ratio between the
advective and reorientation timescales, τa and τr. For τa/τr (cid:29) 1, we expect non-motile microorganisms advected by the flow to
follow the quasi-static reorientation effects described in (b,c); these reorientation effects are strongest near the sinking particle.
this fixed point is an attractive spiral [as in Fig. 2(c) but
with arrows reversed]. This change in the nature of the
convergence of p onto φ indicates the increasing role of
vorticity near the particle equator, but φ remains the at-
tractive fixed point in region I because the fluid has to
expand along the azimuth to accommodate the sinking
sphere in that region. The timescale τI associated with
convergence onto φ in region I is τ−1
2 λ3 = − 3
I = 3
2 F.
three real eigenvalues or a pair of complex eigenvalues
plus a real eigenvalue are separated by two surfaces of
revolution defined by [broken red lines in Figs. 3(b,c)]
r2(θ) = (9 + 4 tan2 θ)/(9 − 4 tan2 θ).
(16)
Furthermore, in the region with complex eigenvalues, the
real eigenvalue, λ3 = −F (r, θ), changes sign from nega-
tive to positive at the plane θ = π/2, which contains the
sinking particle's equator. Physically, the sign change re-
flects the transition of λ3 = φ from being the direction of
fluid expansion to compression as the fluid parcels travel
from the southern to the northern hemisphere. The sur-
faces (16) and θ = π/2 divide the space outside the parti-
cle into three regions I, II and III [Fig. 3(b)]. Region I is
the bottom-half of the entire domain, below the equator
plane θ = π/2 and upstream of the sinking particle. It is
composed of two subregions, Ia and Ib, separated by the
surface (16) [lower broken red line in Fig. 3(b)]. In Ia,
all the eigenvalues are real, in Ib, there is a pair of com-
plex eigenvalues and a positive real eigenvalue λ3 > 0.
In both subregions, the rate of strain dominates over the
rate of rotation and the asymptotic stable direction is
given by the eigenvector λ3 = φ, which always points
along the azimuth. The two subregions differ only in the
manner this asymptotic orientation is approached: in Ia,
the convergence is overdamped [as in Fig. 2(b)] since φ
is an attractive node, in Ib, the convergence is under-
damped with the bacterium spiraling down onto φ since
Region II lies to the side of the particle, in between the
equator plane and the surface (16) and is the only region
in which the rotation rate out-competes the rate of strain.
In this region, λ1,2 are complex and λ3 < 0; physically,
the fluid is being compressed along the azimuth as it is
rolling over the particle surface due to the no-slip bound-
ary conditions. The analysis in Section III A implies that
the rods eventually rotate in the plane orthogonal to the
4β tan2 θ − 9. Thus, rods
azimuth with frequency F/2
orient orthogonal to the vorticity ω and rotate in the
same sense as ω, but the rotation period is longer from
the rotation rate of the fluid [color code in region II in
Fig. 3(b)]. The timescale τII associated with the reori-
entation from pointing along φ to rotating in the plane
perpendicular to φ is given by τ−1
(cid:112)
2λ3 = 3
II = 3
Region III lies downstream of the particle, above the
surface (16). Here, the strain once again dominates over
rotation, but this time the asymptotic direction of rods
in flow is given by the eigenvector λ1 [the white director
field lines in Fig. 3(b)].
Importantly, just behind the
2 F .
0123ρ/R(a)z/R3-3-2-1102(d)Normalized vorticity ║ω║/(U/R)1.400.710210110010-110-2Advection-reorientation timescales τa/τrz/R3-3-2-11020123ρ/R(c)spheres (α=1, γ=0)(b)Rotation frequency / (║ω║/2)100.80.60.40.20123ρ/Rz/R3-3-2-1102region IIIrods (α=∞, γ=1) region Iaregion IIz/R3-3-2-1102disks (α=0, γ=-1) region Iregion IIregion IIIbp(t→∞)p(t→∞)p(t→∞)region Ibregion IIIa0123ρ/RRotation frequency / (║ω║/2)100.80.60.40.2(cid:16)
1 +
(cid:17)
.
2 λ1 = 3
4 F
(cid:112)
9 − 4β tan2 θ
particle, for small colatitudes θ, Eq. (15) clearly pre-
dicts that the stable orientation is approximately the
radial direction λ1 ≈ [1, 0, 0]. The timescale τIII asso-
ciated with the reorientation from rotating in the plane
perpendicular to φ in region II to pointing along λ1 is
τ−1
III = 3
The response of perfect disks (α = 0, γ = −1) is dual to
the case of perfect rods since Aγ=−1 = −E + W = −AT.
For brevity, we only summarize the results, which are
essentially an upside-down version of the responds of
rods [Fig. 3(c)]. Upstream of the sinking particle, disks
tend to be oriented almost tangentially to the particle,
with their symmetry axis pointing in the nearly radial
direction [region I, white director lines in Fig. 3(c)]. To
the side of the particle, disks rotate, with their axis of
symmetry spinning in the r− θ-planes (region II). Down-
stream of the particle (region III), disks preferentially
align with the r − θ planes with their axis of symmetry
pointing along the azimuth. As for rods, the region III
is divided into two subregions: in IIIa, φ is an attractive
spiral, in IIIb it is an attractive node.
The splitting of the fluid flow into the regions shown
in Figs. 3(b,c) is a quasi-static characterization of the
dynamical system (3), with microorganisms held fixed
at a given position in the flow. However, a non-motile
microorganism that is advected by the flow may be sig-
nificantly displaced during the time it takes to achieve
a given asymptotic orientation. To get further insight
into Eq. (3), we compare the two timescales characteriz-
ing the advection and shear-induced reorientation. For
simplicity, we combine the three timescales τI,II,III asso-
ciated with convergence onto the asymptotic solutions
r ∼
to Eq. (3b) into a single reorientation timescale τ−1
F ∼ (cid:107)A(cid:107)2. We estimate the advective timescale τa ∼
R/(cid:107)v(cid:107) at a given position as the time needed to travel
the distance R at the local speed (cid:107)v(cid:107). Fig. 3(d) shows the
ratio τa/τr as a function of the position. In particular, in
the bright oval near the particle τa > τr, indicating that,
in that region, non-motile bacteria advected by the flow
have enough time to orient under the fluid forces in the
manner outlined in Figs. 3(b,c) for immobilized bacteria.
Finally, we note that the eigenvalues of the velocity
gradient A (13) on the stagnation line (θ = 0, π) and the
particle surface (r = 1) have multiplicity greater than
one. In this case, the analysis of Section III A does not
directly apply, yet these special locations are important
for the encounter process of non-motile microorganisms,
which can only approach the sinking particle near the
stagnation line θ = π. In Appendix A 4, we show that on
the upstream stagnation line (θ = π) rods align tangen-
tially to the sinking particle, while on the downstream
stagnation line (θ = 0), rods align vertically. This pic-
ture can be inferred from Fig. 3(b) by taking the limit
ρ → 0. Similarly, disks align tangentially to the particle
surface for θ = π (with axis of symmetry in the verti-
cal direction), while they lie in the r − θ plane for θ = 0.
Therefore, non-motile rods or disks approaching the sink-
7
ing particle along the θ = π stagnation line orient with
their longer dimension tangential to the particle surface.
Furthermore, on the particle surface, shear maintains to
zeroth order the tangential orientation of rods and disks
as they are advected around the sinking particle (Ap-
pendix A 4). This suggests that it is the shorter dimen-
sion of rods and disks that determines their collision with
the particle. However, for any finite size microorganism,
one must step away from the stagnation line and the
particle surface. In the vicinity of these degenerate sets,
the response is captured in Figs. 3(b,c). The asymptotic
orientations rods and disks assume in their respective re-
gions I suggest that the tangential orientation prevails.
However, in regions II and III shear reorients rods and
disks away from the tangential orientation. Given the
size of the regions II and III for rods and disks, this re-
orientation should be stronger for disks, since disks ex-
perience it over a larger part of the particle surface. In
the next section, we use numerical simulations to confirm
this intuition: the collision radius of rods is determined
by their width, not length, whereas for disks the collision
radius is determined by the longest dimension.
IV. NON-MOTILE BACTERIA AND DIATOMS
Understanding the interception of non-motile elon-
gated and oblate microorganisms by a sinking particle
is important for two reasons. First, in its own right, be-
cause many marine microorganisms including many bac-
teria and phytoplankton species are non-motile and come
in a variety of shapes, with bacteria often being spherical
or elongated and phytoplankton being either elongated
(e.g., chains), spherical or disk-lake (e.g., diatoms). Sec-
ond, the non-motile case corresponds to the high sinking
speed limit U/Ub → ∞ for motile bacteria. We predict
drastically different encounter rates for rods and disks:
rods are particularly inefficient at intercepting the sink-
ing particle due shear, which aligns them in the direction
tangential to the particle surface (Section IV A). Con-
versely, disks eventually tumble under shear and explore
their long axis to reach the collector. Experiments on
elongated diatoms support this picture (Section IV B).
A.
Interception of non-motile bacteria
The ballistic interception of non-motile microorgan-
isms by a sinking particle is conceptually identical to
the classical problem of filtration, in which a colloid is
captured by a large collector [9]. Previous works focused
on spherical colloids; through numerical simulations, we
extend these results to nonspherical colloids. Rods and
disks, such as certain species of bacteria or diatoms, have
drastically different effective collision radii. Due to shear-
induced reorientation, the collision radius for a rod is
determined by its width rather than length, while disks
explore their full size to intercept the collector.
8
dius is rc ≈ 1.4(lb/2); the additional 40% arise due to
the squeezing of the streamlines near the collector. Had
we traced the streamline all the way to z → −∞, the
prefactor would change from 1.4 to 1.2 [9]. In general,
rc depends very weakly on the size of spherical bacteria
lb/R and the formula rc ≈ 1.4(lb/2) works very well for
the bacterial sizes in the range 0 < lb/R < 1/10. Fi-
nally, for spherical colloids, the effective collision radius
and the encounter efficiency [Eq. (7)] are related as
ηspheres = (rc/R)2.
(18)
We now turn to nonspherical organisms, for which the
orientational dynamics can no longer be neglected.
As nonspherical microorganisms follow the stream-
lines, they can intercept the sinking particle using either
their shorter or longer dimension. Two extreme scenar-
ios are possible: an organism always aligns its longer
side tangentially or perpendicular to the particle, which
modifies its collision efficiency by a factor of α2. As-
suming negligible rotational diffusion, shear impacts the
orientation of nonspherical organisms only through the
aspect ratio α. This follows from nondimensionalizing
Eqs. (3) with Ub = 0 in terms of the particle radius R
and timescale R/U . However, while the organism size
lb does not directly affect the dynamics, it determines
the interception criterion: we take the sinking particle to
be a perfect absorber and stop simulations if any part
of the rod or disk touches the particle (Appendix A 5).
We now systematically vary α and lb/R and measure the
encounter probability and typical interception location.
The impact of the microorganism aspect ratio α and
its relative size lb/R on the encounter problem is sum-
marized in Figs. 4(b -- d); lb denotes the longer dimen-
sion - length for elongated organisms, width for oblate
ones. Varying α [Fig. 4(b)] at fixed R/lb = 10, shows
that the collision radius of elongated microorganisms is
determined by their width, not length (Movie 1). This
is evident from the variation of P (ρ), the interception
probability for an initial position at distance ρ from
the centerline: P (ρ) decreases sharply from one once
ρ > rc/α. The probability tail between rc/α < ρ < rc
indicates that rods occasionally reorient and use their
length to intercept the particle (Movie 2). However, shear
largely suppresses this effect, see the light blue broken
line in Fig. 4(b), which represents trajectories without
the shear-induced reorientation (parallel transport). In
contrast to elongated microorganisms, oblate organisms
(disks) utilize their full size to intercept the particle -
P (ρ) drops sharply from one to zero near rc. To see the
impact of varying the relative size R/lb, we fixed two as-
pect ratios, α = 10 [Fig. 4(c)] and α = 0.1 [Fig. 4(d)]
and computed P (ρ) as well as the distribution ξ(θ) for
R/lb = 10, 100. We observe that, as the colloid gets
smaller (or the sinking particle gets larger), the effects
described above become more pronounced, in the sense
that for rods the probability tail between rc/α < ρ < rc
shrinks, whereas for disks P (ρ) approaches a step func-
tion with jump at rc. Therefore, the formula (18) for the
FIG. 4: Generalizing the classical ballistic interception prob-
lem to nonspherical colloids reveals that shear orients non-
motile elongated microorganisms tangentially to the sinking
particle surface, while oblate organisms tumble near the par-
ticle. (a) For spherical organisms, the effective collision radius
rc is defined by the critical streamline beyond which no inter-
ception occurs (white line). (b) Encounter probability P (ρ)
as a function of distance ρ from the origin in the initial plane
z = −6R for different aspect ratios α. For spheres and disks,
P (ρ) drops from 1 to 0 near rc, whereas for rods the drop oc-
curs near rc/α. (c) The probability P (ρ) for rods (α = 10) for
two different sizes R/lb = 10, 100 shows that the probability
tail between rc/α and rc vanishes as the sinking particle size
grows (or the rod becomes smaller). The tail arises to due rare
tumbling events to the leeward side of the particle (Movie 2),
as can be seen from the distribution of the interception co-
latitudes ξ(θ) (inset). (d) Conversely, P (ρ) for α = 0.1 for
R/lb = 10, 100 demonstrates that disks tumble very often -
as the particle grows, P (ρ) approaches a step function at rc.
First, we briefly review the classical interception of
small non-motile spherical beads by a large spherical col-
lector. The reorientation of beads under the flow does not
affect the interception problem, which reduces to identi-
fying the streamline of closest approach [Fig. 4(a)]. For
the Stokes flow (2), the stream-function is given by
(R/r)3(cid:17)
sin2 θ.
(17)
U r2(cid:16)
ψ(r, θ) =
1
2
1 − 3
2
R/r +
1
2
The streamlines determined by Eq. (17) have the fore --
aft symmetry, which implies that the critical streamline
separating captures from misses is defined by the point
r = R + lb/2 and θ = π/2. Tracing the streamline up-
stream from this point to z = −6R, where we start the
simulations, defines the effective collision radius rc. For
a spherical bacterium with lb/R = 1/10, the collision ra-
(a)(b)R/lb=102-6-40z/R-202ρ/R00.7501.5-5.5-1.50lb/2-6rc~1.4 (lb/2)lb(c)(d)α=10, γ=0.98R/lb=10R/lb=100α=0.1, γ=-0.98R/lb=10R/lb=10010-210-445°135°180°Colatitude θ90°1ξ(θ)20×10-2rc/αα=1α=2α=10α=.5α=.1lblbα=10, no shear01.5Encounter prob. P(ρ)ξ(θ)45°135°180°Colatitude θ90°01.5Encounter prob. P(ρ)01.5Encounter prob. P(ρ)00.40.8ρ/lbrc00.40.8ρ/lbrc00.40.8ρ/lbrcRθ9
FIG. 5: Experiments with non-motile elongated diatom cells (Phaeodactylum tricornutum) are consistent with the predictions
of the model [Eq. (3)] that rods maintain tangential orientation as they are advected around a sinking particle (Section IV A).
(a) Minimum intensity projection obtained by phase contrast microscopy shows the streamlines of the suspended diatoms
around the alginate particle (Movie 3). (b) The experimental ensemble of diatom positions (dots) and the sine of the angle
the diatoms make with the flow direction (see colorbar). (c) The corresponding ensemble obtained from simulations, in which
we mimic the same information loss as in the experiment: the plot represents rods lying in the focal plane and we dismiss
rods with a significant out-of plane component, see text for more details. As predicted in Section A 4, we observe that rods
approaching the particle near the upstream stagnation line orient tangentially to the particle [yellow region below the particle
in (b,c)], while rods that leave the particle near the downstream stagnation line point nearly radially (blue region above the
particle). As they travel near the particle surface, rods tend to maintain tangential orientation but also occasionally tumble.
encounter efficiency by spherical colloids is replaced by
ηrods = [rc/(αR)]2 = ηspheres/α2,
ηdisks = ηspheres (19)
in the case of (small R/lb (cid:29) 10) nonspherical colloids,
where rods means α (cid:29) 1 and disks α (cid:28) 1.
Different interception efficiency for rods but not disks
as compared to spherical colloids is consistent with the
analytical arguments presented in Section III B. Initially,
shear aligns rods and disks tangentially to the sinking
particle surface as they approach it along the stagnation
line [regions I in Figs. 3(b,c)]. As they slide near the
particle, both rods and disks experience shear that tries
to reorient them away from the tangential configuration,
potentially increasing their chance to intercept the par-
ticle [region II in Fig. 3(b) for rods and regions II and
IIIa in Fig. 3(c) for disks]. However, disks are exposed to
this reorienting effect over a larger region than rods, sug-
gesting that disks complete this reorientation while near
the particle, whereas rods orient radially only when they
are too far behind the particle (Movie 1). Occasional in-
terception by rods caused by the reorientation in region
II is responsible for the small probability tail in P (ρ) in
Figs. 4(b,c) (Movie 2). In summary, simulations confirm
the intuition based on analytical arguments: the colli-
sion cross-section for rods is determined by their shorter
dimension whereas the opposite is true for disks.
B. Experiments with elongated diatom cells
Selected experiments with non-motile diatom cells con-
firm the predictions of the model (3) discussed in Sec-
tions IV A, A 4 in the case of the non-motile rods (Fig. 5
and Movie 3). We ran a suspension of the non-
motile, elongated diatom cells Phaeodactylum tricornu-
tum (strain CCMP2561) in sea water at a mean flow
velocity of 168 µm s−1 through a microfluidic channel
with a calcium-alginate spherical particle held fixed in
the middle by the channel walls [Fig. 5(a)]; see Ap-
pendix A 6 for more details on the experimental proto-
col. The particle size was R = 566 µm, the average di-
atom length lb = 21.2 µm and their average aspect ratio
α = 6.8. Rather than directly estimating the encounter
rate, which proved difficult due to the challenge of imag-
ing in the immediate vicinity of the particle, we used
image analysis to quantify the orientation that the di-
atoms assume in the vicinity of the particle in the chan-
nel mid-plane. Using this approach, we extracted from 50
consecutive frames an ensemble of diatom positions and
orientations [Fig. 5(b)]. Since the shear-induced reorien-
tation effects are strongest near the particle [Fig. 3(d)],
we focus on diatoms that are at distance R < r < 2R
away from the particle center. Importantly, since we only
image the focal plane, this ensemble is skewed towards
diatoms moving in the focal plane and also oriented in
that plane. For this reason, to compare the experimen-
tally determined orientations with those predicted by the
model (3), we run additional numerical simulations to
mimic the same information loss as in the experiments.
Specifically, we simulate a front of uniformly distributed
and randomly oriented elongated non-motile rods (with
R/lb = 21.2 and α = 6.8), as in the previous section. We
focus on trajectories lying in the particle mid-plane (as
in the imaged region) and extract the rod orientations at
positions along the streamlines corresponding to equal
time intervals. We reject orientations that have the out-
(a)(a)(b)experiment(c)theory100.5Sine of the angle with verticalUx/R0-22-11210z/R-2-1x/R0-22-11210z/R-2-1experiment10
FIG. 6: The shear-shape coupling significantly impacts the encounter rate between motile bacteria and sinking particles
N /n (a), encounter efficiency η (b) and mean
and the typical interception location on the particle. Encounter rate kernel
interception colatitude (cid:104)θ(cid:105) (c) as a function of the sinking speed relative to the bacterial swimming speed U/Ub for different
bacterial aspect ratios α. The continuous lines represent the ballistic model (3) while the broken lines denote the quasi-ballistic
model with rotational diffusion (20). In the ballistic case, motility, elongation and shear enhance the encounter rate about
twofold for slowly sinking particles as compared to the case with shear-induced reorientation switched off (purple and green
lines). However, for intermediate to fast sinking particles, the encounter rate falls orders of magnitude below the value set by
the interception of non-motile rods. On the particle, elongated motile bacteria attach preferentially to its leeward side.
of-plane component larger than sin(30°) = 0.5, to mimic
the information loss of diatoms that point out of the fo-
cal plane in experiments; the results are robust against
variation in this threshold (Fig. A.1).
[Figs. 5(b,c)].
The experimentally determined orientations of diatoms
agree very well with the numerical results for elon-
gated ellipsoids with the same geometrical characteris-
tics
In particular, as predicted in Ap-
pendix A 4, diatoms approaching the particle near the
upstream stagnation line orient tangentially to the par-
ticle surface, while diatoms departing from the particle
near the downstream stagnation line point nearly radially
[yellow vs. blue regions in Figs. 5(b,c)]. Close to the par-
ticle surface, diatoms tend to maintain tangential orienta-
tion but can also occasionally tumble. Tumbling happens
most often when diatoms are to the leeward side of the
particle, which is consistent with the action of shear de-
picted in region II in Fig. 3(b), where vorticity dominates
over straining and tries to spin rods in the plane of the
picture. In summary, these experiments validate detailed
aspects of our model for non-motile microorganisms and
demonstrate that the effect of shear -- shape coupling can
be substantial for realistic marine microorganisms.
V. MOTILE BACTERIA
The encounter rate between motile elongated bacte-
ria and sinking particles in the ballistic regime depends
strongly on the particle sinking speed relative to the bac-
terial swimming speed (Section V A). For slow sinking
particles, shear increases the encounter rate more than
twofold and leads to preferential attachment of bacte-
ria to the leeward side of the particle. However, as the
sinking speed increases, shear decreases the encounter
rate, orders of magnitude below the rate of non-motile
organisms. These mechanisms of hydrodynamic focusing
and screening are rationalized at the level of individual
bacterial trajectories (Section V B) in terms of the quasi
static picture derived in Section III. Finally, to connect
with the diffusive description of the encounter process,
we introduce rotational diffusion to quantify how vari-
ous stochastic mechanisms, such as Brownian motion or
run-and-tumble reorientation, influence the above ballis-
tic description (Section V C).
A. Encounter rates for motile bacteria
Nondimensionalization of the ballistic model (3) in
terms of the particle radius R and the time scale R/U
derived from the sinking speed U shows that the only two
dynamically relevant variables are the ratio of the sink-
ing to swimming speeds U/Ub and the bacterial aspect
ratio α. The bacterial size lb and particle size R enter
the problem through the interception condition but oth-
erwise they do not affect the bacterial trajectories, except
for the time it takes to execute them. We assume the par-
ticle is a perfect absorber and stop the simulations either
if any part of the bacterium touches the particle (inter-
ception) or the bacterium ends up far behind the particle.
In this section, we fix R/lb = 10, scan velocities in the
range U/Ub > 1 (sinking speed greater than swimming
speed) and consider several aspect ratios α (Fig. 6).
For hypothetical spherical or oblate motile bacte-
N /n depends weakly
ria [22], the encounter rate kernel
on the sinking velocity U/Ub and the encounter efficiency
η decays monotonically with U/Ub; for spherical swim-
mers, η is close to the values obtained with the reorien-
tation by shear switched off [dark blue and green lines in
Figs. 6(a,b)]. However, for elongated swimmers, η varies
strongly with U/Ub [Fig. 6(a,b)]. For slowly sinking par-
101102Flow speed U/Ub100103102101Encounter rate kernel N/n/(lb3τb-1).10010-210-4Encounter efficiency η0°90°180°Mean colatitude ‹θ›(c)(b)(a)downstreamupstreamshear OFF1asp. ratio α1.755100.1ηspheresEq. (18)101102Flow speed U/Ub100101102Flow speed U/Ub10011
FIG. 7: Motile elongated bacteria preferentially attach to the leeward side of a sinking particle due to hydrodynamic screening
and focusing upstream and downstream of the particle. Encounter probabilities P (x) as a function of the position x [Eq. (4)]
for perfect spheres (a), moderately elongated swimmers (b) and perfect rods (c). The sinking to swimming speed ratio is fixed
at U/Ub = 3. (d -- f) Histograms of the interception colatitudes θ for initial positions in the whole domain shown in (a -- c) show
a transition from a nearly uniform coverage of the particle by spherical swimmers (d) to preferential leeward attachment for
motile rods (f). (g-i) Representative swimming trajectories (left) and successful initial orientations (right) for rod-like bacteria
starting from the initial positions indicated in (c) illustrate the hydrodynamic screening (g) and focusing (h,i).
ticles (1 < U/Ub < 2), η ∼ 2 − 3, implying that the
particle collects bacteria from the volume of water two-
three times bigger than the geometric cylinder the par-
ticle swipes as it sinks. Furthermore, elongated bacteria
intercept the particle to the leeward side [Fig. 6(c)]. In-
terestingly, as the sinking speed increases, the encounter
rates drop very rapidly:
in the the velocity window
10 < U/Ub < 100, the encounter rate of elongated swim-
mers (α ≥ 5) can be orders of magnitude below the value
set by the non-motile rods. We next rationalize these
encounter rate enhancement and decrease using the con-
cepts of hydrodynamic focusing and screening.
B. Hydrodynamic focusing and screening
The strong dependence of the encounter efficiency
on the particle sinking speed for elongated bacte-
ria [Fig. 6(b)] is a consequence of hydrodynamic focus-
ing and screening. These phenomena are illustrated in
Fig. 7, where we compare three swimmers with different
aspect ratios; the sinking speed is fixed at U/Ub = 3.
Figs. 7(a -- c) show the encounter probabilities P (x) for
a bacterium starting at x anywhere inside the indicated
domain (not just the plane z = −6R) with head point-
ing in a random direction [Eq. (4)]. Since U/Ub > 1, in
all three cases there is a cone-like surface of revolution
that separates the accessible [P (ρ) > 0] and inaccessi-
ble initial positions - if the bacteria start too far away,
they cannot reach the particle. The distribution of P (x)
inside the accessibility region vary strongly with α. For
spheres, P (x) is concentrated below the particle, near
the ρ = 0 stagnation line and decays monotonically to
zero with ρ reaching the accessibility horizon [Fig. 7(a)].
For somewhat elongated swimmers α = 1.75, the initial
positions below the particle become less likely to result
in an interception and P (x) starts to concentrate near
the edge of the accessible region, which also reaches fur-
ther out [Fig. 7(b)]. For perfect swimmers α = ∞, the
region ρ ≈ 0 is now almost entirely shielded, with P (x)
exhibiting a clear high-probability belt at the edge of the
accessible region [Fig. 7(c)]. Far below the particle, the
belt slope approaches ∼ U/Ub. Considering the distribu-
tion of the interception locations ξ(θ) for initial positions
anywhere in the domains shown in Figs. 7(a -- c), elon-
gated swimmers show preferential leeward attachment,
with the vicinity of the 'north pole' being the most likely
location [Figs. 7(d -- f)]. We now rationalize the shape of
the distributions P (x) and ξ(θ) at the level of individual
swimming trajectories by evoking the quasi-static picture
discussed in Section III B and shown in Fig. 3.
At the level of individual swimming trajectories, the
probability P (x) for spherical swimmers [Fig. 7(a)] is
realized by trajectories that correspond to swimmers
initially located below the particle and pointing up-
wards. However, this intuitive strategy is not avail-
(g)(h)-2(i)0222-20z/Ry/Rx/R01-11-10201040222-20-41.95-2.05-2.1.15.05.05-.050-3.95-4.05-4z/R1.851.751.8.05-.050y/R.05-.05044.053.95x/R.01.06-.04(b)(e)1.50×10-2α=1.75, γ~0.5ξ(θ)1.00.50°90°180°Colatitude θ4-6-4-220z/R0246ρ/R(c)(f)60×10-2(i)(g)(h)α=∞, γ=14-6-4-220z/REncounter probability P(x)0.400.20246ρ/Rξ(θ)3~U/Ub0°90°180°Colatitude θ(a)(d)40×10-3α=1, γ=0ξ(θ)24-6-4-220z/R0246ρ/R0°90°180°Colatitude θRθ-2-2012
FIG. 8: Encounter probability P (ρ) (a,b) and distribution of interception colatitudes ξ(θ) (c,d) for bacteria starting at the
plane z = −6R with a random initial orientation for different relative sinking speeds U/Ub and aspect ratios α; (a,c) shows
the results of numerical simulations of the ballistic model (3) and (b,d) the stochastic model (20). In the parameter range
considered, rotational diffusion mainly affects elongated swimmers (bottom panels), for which it decreases the impact of the
hydrodynamics focusing at low sinking speeds but also ameliorate the hydrodynamic screening at higher sinking speeds.
able for elongated swimmers because of hydrodynamic
screening [Fig. 7(b,c)]. Recall that, below the particle,
shear tends to align rods along the azimuth [region I in
Fig. 3(b)]. This shear-induced reorientation coupled with
forward motility implies that rod-like swimmers get reori-
ented and swim away as they approach the sinking parti-
cle from below [Fig. 7(g), Movie 4]. For the same reason,
it is very unlikely that elongated swimmers attach to the
front of the particle, which explains the small values of
ξ(θ) for colatitudes θ > 90° [Fig. 7(f)].
Instead, suc-
cessful interceptions for elongated bacteria must follow a
different strategy [Figs. 7(h,i)]. To avoid the screening,
elongated swimmers must start on the belt far away from
the centerline of the sinking particle, on the edge of the
accessibility horizon. Furthermore, their initial orienta-
tions have the be roughly horizontal, pointing towards
the centerline [Figs. 7(h)]. Such initial conditions allow
the bacteria to avoid the screening region I of Fig. 3(b)
and explore the shear-induced radial reorientation in re-
gion III. This hydrodynamic focusing then leads to pref-
erential leeward attachment (Movie 5).
The mechanisms of hydrodynamic focusing and screen-
ing described above rationalize the strong dependence of
the encounter efficiency η on the particle sinking speed
U/Ub presented in [Fig. 6(b)]. For slowly sinking speeds,
both mechanisms are present. However, the high proba-
bility belt at the edge of the accessibility horizon for elon-
gated swimmers extends a large volume and hence many
swimmers can utilize the focusing effect, which explains
why η > 1 in that flow range. However, as U/Ub in-
creases, the high probability belt moves closer to the cen-
ter line since its diameter scales as Ub/U . This reduces
the accessible volume of water at the rate at least ∼ U−2.
Furthermore, as the belt shrinks in diameter, it enters the
region of hydrodynamic screening and eventually disap-
pears [Fig. 8(a)]. Thus, in the range 10 < U/Ub < 100,
only screening persists, which explains the very small
values of η in that range. Only for swimming speeds
U/Ub > 100, η rises again, until it starts to recover the
limit set by the interception rate of non-motile rods.
C.
Impact of rotational diffusion
In the purely ballistic picture of the encounter process
outlined in the two previous sections, shear is the only
factor responsible for microorganism reorientation.
In
reality, bacteria experience Brownian rotational diffusion
as well as perform run-and-tumble or run-and-reverse dy-
namics. The combination of these stochastic mechanisms
likely interferes with the shear-induced reorientation in a
complex manner. As a first step to systematically study
the impact of these additional mechanisms, we introduce
a single rotational diffusion term to Eq. (3)
x = Ubp + v,
p = (I − ppT)[(γE + W )p +
(20a)
(20b)
2Drξ],
(cid:112)
where Dr is the cell's effective rotational diffusivity and
ξ is a delta-correlated 3D white noise with zero mean.
We express the diffusive timescale τd = D−1
in terms of
the time τb = lb/Ub needed for a bacterium to travel the
distance equal to its bodylength.
r
To study the impact of rotational diffusion on the en-
counter rates and attachment location, we fixed the dif-
fusive timescale at τd/τb = 100, which corresponds to the
(a)(c)(b)α=1α=10P(ρ)10010-110-2P(ρ)10010-110-210-210-1100101ρ/R10-210-1 10010110-3U/Ub=100U/Ub=1.25U/Ub=2U/Ub=5U/Ub=200(d)ξ(θ)10-110-210-30°90°180°Colatitude θ×10-3804ξ(θ)α=1α=1010010-110-210010-110-2ρ/R10-210-1 10010110-3α=1α=10diffusion ON, τd/τb=100×10-380410-210-11001010°90°180°10-210-11001010°90°180°0°90°180°Colatitude θ10-110-210-3α=1α=10diffusion ON, τd/τb=10010-210-11001010°90°180°10-210-11001010°90°180°P(ρ)10010-110-210010-110-2×10-3402ξ(θ)×10-3402U/Ub=30U/Ub=1.25U/Ub=5U/Ub=10U/Ub=200U/Ub=1.25U/Ub=5U/Ub=10U/Ub=30U/Ub=200α=0.1α=0.1α=0.1α=0.1diffusion OFFdiffusion OFFU/Ub=30U/Ub=1.25U/Ub=5U/Ub=10U/Ub=200U/Ub=30U/Ub=1.25U/Ub=5U/Ub=10U/Ub=200U/Ub=1.25U/Ub=5U/Ub=10U/Ub=30U/Ub=200U/Ub=1.25U/Ub=5U/Ub=10U/Ub=30U/Ub=200U/Ub=30U/Ub=1.25U/Ub=5U/Ub=10U/Ub=200U/Ub=30U/Ub=1.25U/Ub=5U/Ub=10U/Ub=200U/Ub=100U/Ub=1.25U/Ub=2U/Ub=5U/Ub=200U/Ub=100U/Ub=1.25U/Ub=2U/Ub=5U/Ub=200U/Ub=100U/Ub=1.25U/Ub=2U/Ub=5U/Ub=200typical time-scale set by the run-and-tumble motility [8].
We then repeated the scans according to the same proto-
col as in Section V A, with R/lb = 10 and U/Ub > 1 (the
broken lines in Fig. 6). We find that diffusion has lit-
tle effect on the encounter rates and attachment location
for spherical swimmers. However, for elongated swim-
mers, diffusion decreases the impact of hydrodynamics
focusing at low sinking speeds but also decreases the hy-
drodynamic screening at higher sinking speeds.
The uniformizing impact of diffusion is studied in more
detail in Fig. 8, where we consider the interception proba-
bility P (ρ) for a bacterium starting at the plane z = −6R
with random orientation [Figs. 8(a,b)] as well as the
corresponding distribution of the interception locations
ξ(θ) [Figs. 8(c,d)]. We compare side to side the cases
without [Figs. 8(a,c)] and with diffusion [Figs. 8(b,d)] for
oblate, spherical and elongated swimmers (top, middle
and bottom rows, respectively). The accessibility region
[defined as P (ρ) > 0] shrinks under diffusion, because the
now erratic motion of bacteria takes longer to reach the
particle. and therefore, the bacteria must start closer to
the particle to be able to catch it. Within the accessibil-
ity region, the distribution P (ρ) for oblate and spherical
swimmers is nearly unaffected by diffusion [Figs. 8(a,b),
top and middle rows] and so is ξ(θ) [Figs. 8(c,d), top
and middle rows]. However, for elongated swimmers,
diffusion decreases the size of the high probability belt
near the edge of the accessibility region but also raises
the probability of interception for initial conditions di-
rectly below the sinking particle [Figs. 8(a,b), bottom
row]. As a consequence, diffusive elongated swimmers
have nonnegligible probability of attaching to the front
of the sinking particle [Figs. 8(c,d), bottom rows].
13
FIG. 9: Given a fixed volume Vb of a non-motile microorgan-
ism, what shape minimizes or maximizes the encounter effi-
ciency η with a sinking spherical particle of volume V ? Plot-
ting η vs the volumetric ratio Vb/V reveals that elongation
reduces the encounter rate with large particles (Vb/V → 0),
making rods the optimal shape for avoiding sinking particles,
at least as long as rotational diffusion is negligible. Con-
versely, flattening makes a non-motile microorganism partic-
ularly efficient at intersecting the particle. This different be-
havior of rods and disks is a consequence of fluid shear, which
aligns rods tangentially to the sinking particle surface with
rare tumbling events, but induces frequent tumbling in the
orientation of disks. As the sinking particle becomes large,
Vb/V → 0, the efficiencies for rods and disks approach the
exact expression η = 1.42[Vb/(V α)]2/3, which follows from
Eqs. (19) after assuming rods and disks can be represented as
prolate and oblate ellipsoids with aspect ratio α. As the vol-
ume ratio Vb/V grows, the difference between rods and disks
decreases due to increasing tumbling of rods.
VI. DISCUSSION
In this work, we have studied the ballistic limit of the
encounter process between microorganisms and sinking
particles, with focus on the reorienting effect of shear
induced by the sinking particle. Analytical and numeri-
cal calculations as well as selected experiments show that
the shear -- shape coupling acting on a microorganism im-
pacts population-level observables, such as the encounter
rate with sinking particles or the typical attachment loca-
tion on the particle. For the Stokes flow around a spheri-
cal sinking particle, rods and disks break the fore-aft sym-
metry of the flow streamlines, in stark contrast to the be-
havior of spherical colloids (Fig. 3). This shape-induced
symmetry breaking affects the encounter rates (Figs. 4
and 6) through mechanisms we have characterized as hy-
drodynamic focusing and screening (Fig. 7). Below, we
first rephrase these results as solutions to the optimiza-
tion problem: should a microorganism be elongated or
flat, to maximize or minimize the encounter rate with a
large moving sphere? Subsequently, we discuss the bio-
physical consequences of our mechanistic description of
the encounter process in the marine environment.
From the perspective of evolution, there are many
contexts in which microorganisms may seek to maxi-
mize or minimize their encounters with moving objects,
including encountering sinking resources [8] or symbi-
otic partners [23], and avoiding predators [24]. At the
same time these, microorganisms are likely to have other
constraints on their volume, such as growth maximiza-
tion and genome size. For non-motile microorganisms
with negligible rotational diffusion (Section IV A), we
have seen that shear tends to orient rods tangentially
to the sinking particle surface as these move around
the particle, whereas disk-shaped microorganisms tum-
ble, which makes their longer dimension available for in-
terception (Section IV A and Fig. 4). As a consequence,
rods have their encounter efficiencies decreased by a fac-
tor equal to the square of their aspect ratio compared to
spherical colloids with diameter equal to the rod length.
Conversely, disks have the same efficiency as spheres with
diameter equal to the disk diameter [Eqs. (19)].
For non-motile bacteria, over a broad range of ratios of
cell volume relative to particle volume (Vb/V ), disks are
the most efficient shape to intercept a sinking particle,
while rods are the least efficient (Fig. 9). The contribu-
Encounter efficiency η10-210-410-61.42[Vb/(Vα)]2/3α=1α=0.1α=10sideviewtopview10-1010-810-410-6 Volumetric ratio Vb/Vtion of the occasional tumbling of rods grows with Vb/V ,
decreasing the difference between the efficiencies of rods
and disks as cell volume gets larger; rods remain less effi-
cient than disks but catch up with spheres. Note that we
cannot increase the volumetric ratio further without vi-
olating the approximations used in the Jeffrey equation,
since the microorganism size becomes comparable with
the sinking particle. Conversely, as the sinking particle
grows, Vb/V → 0, the encounter efficiencies approach
1.42[Vb/(V α)]2/3 (broken lines in Fig. 9), which follows
directly from Eqs. (19) and reflects the fact that, as the
sinking particle grows, rods cease to tumble while disks
always tumble [Fig. 4(c,d)]. In this limiting regime, the
efficiency ratio between disks and rods is (αrod/αdisk)2/3.
For example, for rods with αrod = 10 and disks with
αdisk = 0.1, disks will be about 20 times more efficient
than rods at intercepting large sinking particles.
For motile microorganisms, encounter rates are con-
trolled by hydrodynamic focusing and screening, which
develop from the influence of shear on the direction of
swimming. Here we still consider the ballistic regime,
with negligible rotational diffusion.
In addition, the
shear -- shape coupling for motile cells leads to the sen-
sitive dependence of the encounter rate on the parti-
cle sinking speed relative to the bacterial swimming
speed (Section V A and Fig. 6). Hydrodynamic screen-
ing and focusing describe the effect on elongated motile
microorganisms of the shear upstream and downstream
of the particle, respectively. In hydrodynamic screening,
the shear upstream of the particle aligns rods tangen-
tially to the particle surface (as with non-motile cells),
resulting in cells swimming away from the particle. For
hydrodynamic focusing, rotation from shear in the down-
stream half of the particle turns the swimming rods to-
wards the particle. For slowly sinking particles, focus-
ing dominates and enhances the encounter rate (Fig. 7),
while at sinking speed significantly higher than the cell
swimming speed, screening dominates and the encounter
rate drops far below that of non-motile rods.
In con-
trast to elongated organisms, hypothetical spherical and
disk-shaped swimmers respond monotonically to changes
in the sinking speed (Fig. 6), a consequence of disk-
shaped swimmers experiencing hydrodynamic focusing
upstream, not downstream, of the particle. For sinking
speeds more than twice the swimming speed, this results
in disk-shaped microorganisms encountering particles at
a higher rate than elongated microorganisms.
From an evolutionary perspective, this analysis sug-
gests that adopting a disk shape would be optimal for
microorganisms (whether non-motile or motile) in order
to maximize particle encounter rates at a broad range of
sizes and sinking speeds. Of course, this neglects other
evolutionary pressures on morphology. However, for par-
ticles sinking at speeds close to the organism swimming
speeds (i.e., relatively slowly), elongation increases the
encounter rates up to twice that of motile disks. Cou-
pled with the large reduction in encounter rates for more
rapidly sinking particles, elongation can be viewed as bi-
14
asing ballistically swimming organisms strongly towards
slowly sinking particles. This could be subject to se-
lective pressures for microorganisms in situations where
optimal growth occurs near the surface, and therefore
rapidly sinking particles are better avoided.
For the specific case of marine bacteria encounter-
ing sinking marine particles, it is necessary to connect
the ballistic description of the encounter process with
the classical approach based on approximating bacterial
motility as a diffusive process [7, 8, 25]. Marine bacteria
are subject to various sources of random reorientation,
from Brownian rotational diffusion to self-generated run-
and-tumble or run-and-reverse motility, where segments
of straight swimming are interrupted by randomization
of the swimming direction. As a consequence, on scales
larger than the bacterial run length and timescales longer
than the reorientation time, bacterial motility can be ef-
fectively characterized as a diffusive process [7, 8, 25] -
this is a general feature of superimposing a large num-
ber of uncorrelated random segments [26]. In this limit,
relevant to large sinking particles, the encounter rate is
proportional to the bacterial effective diffusion coefficient
and the Sherwood number, a flow-induced enhancement
factor [25]. Additionally, while the shear-induced reori-
entation in the diffusive limit can be neglected in certain
regimes [12], precise quantification of its impact on the
encounter rate and attachment locations across a wide
range of particle speeds and sizes might require kinetic
theory approach [26 -- 28]. However, as the particle be-
comes smaller or the sinking speed increases, the system
becomes ballistic and the diffusive approximation overes-
timates the encounter rate. The reason for this overesti-
mation comes from the fact that the encounter probabil-
ity for a bacterium at distance r from the particle decays
as r−1 in the diffusive regime, and as r−2 in the ballistic
regime [29], at least for stationary particles and without
shear. Since hydrodynamic interactions can significantly
bend bacterial trajectories [14, 30, 31], the need to go
beyond arguments based on straight-line swimming mo-
tivated the above study of the pure ballistic limit. For
the marine application, we consider ballistic motile bac-
teria supplemented by rotational diffusion (Section V C),
which we have shown reduces the strength of hydrody-
namic screening on rod-shaped swimming cells.
We consider two major classes [8]: motile elongated
bacteria and non-motile spherical bacteria. For motile
elongated bacteria, we evaluate the predictions of our
model for cells of length lb = 2 µm, swimming speed
Ub = 50 µm s−1 and aspect ratio α = 3.3. For non-
motile spherical bacteria, we choose a diameter lb =
1 µm. These represent typical characteristics of motile
copiotrophic bacteria that actively seek and engage ma-
rine particles, and more oligotrophic non-motile bacteria
which may nevertheless encounter and stick to particles.
For these representative marine bacteria, as well as the
equivalent motile bacteria without the influence of shear,
the corresponding encounter efficiency η [Fig. 10(a -- c)]
and mean interception colatitude (cid:104)θ(cid:105) [Fig. 10(d -- f)] have
15
FIG. 10: To quantify the impact of shear and elongation on the encounter rates and interception locations between bacteria
and sinking particles in the ocean, we focus on realistic model parameters and compare three cases: motile elongated bacteria
of length lb = 2 µm, swimming speed Ub = 50 µm s−1 and aspect ratio α = 3.3 (a,d); the same bacteria, but with the shear-
induced reorientation switched off (b,e); and non-motile spherical bacteria of diameter lb = 1 µm (c,f). In these three cases,
we compute the encounter efficiency η (a -- c) and mean interception colatitude (cid:104)θ(cid:105) (d -- f) as a function of the sinking particle
speed U and radius R. We consider a range of sinking particle sizes (R ∼ 3 µm − 1mm) that covers the most abundant marine
particles [11]. In the simulations, the rotational diffusion coefficient for motile swimmers was set to Dr = 0.25s−1, while the
translational diffusion coefficient of the non-motile spheres was set to Dt = 0.43 µm2 s−1. For such parameters, the panels
capture the ballistic and diffusive regimes, as well as the in-between quasi-ballistic regime.
been computed as a function of the sinking particle speed
U and particle radius R. The range of sinking speeds
we consider is 60 µm s−1 − 5 mm s−1. The range of parti-
cles sizes, 3 µm−1 mm, covers the most abundant marine
sinking particles [11]. For elongated bacteria, randomiza-
tion of orientation is effectively represented by a single ro-
tational diffusion coefficient Dr = 0.25 s−1; the diffusive
timescale (∼ 4 s) gives the run length of about 200 µm,
typical of marine bacteria [8]. For such a run length, the
range of particle sizes spans the ballistic and diffusive
regimes as well as the intermediate transition range. Fi-
nally, the translational diffusion of the non-motile spher-
ical bacteria was set to Dt = 0.43 µm2 s−1, which repre-
sents Brownian motion of a micron-sized sphere at room
temperature [32]. Under these conditions, accounting for
the shear reorientation of motile marine bacteria sub-
stantially reduces the encounter efficiency for small fast-
sinking particles and alters the location of encounters for
slow-sinking particles below 100 microns in radius.
In
contrast, the diffusive spherical swimmers show weak de-
pendence of encounter efficiency on sinking speed, but
much greater sensitivity to particle size.
Since the range of particle sizes considered in Fig. 10
captures the ballistic -- diffusive transition,
the stan-
dard computation based on a diffusive analogy only
matches the encounter rates of marine bacteria (neglect-
ing shear) for the largest particles with radii approaching
1 mm (Fig. A.2). For bacteria with higher rotational dif-
fusivity, this occurs for smaller particle sizes. As the par-
ticles get smaller, the diffusion-based calculation starts
to overestimate the encounter rate - in the ballistic limit,
with particle sizes reaching tens of microns, the two de-
scriptions can differ by more than two orders of magni-
tude (Fig. A.2). We now describe in detail the encounter
process in the intermediate quasi-ballistic regime, high-
lighting the role of bacterial motility and fluid shear.
Factoring in shear interactions, motile bacteria en-
counter sinking particles at a rate one or two orders of
magnitude [Fig. 11(a)] greater than non-motile bacteria.
This motility-based enhancement factor is smaller by one
or two orders of magnitude (depending on particle size)
as compared to what would be predicted by the fully dif-
fusive model. The exception when motility decreases the
chances of interception (ηmotile
shear ON/ηnon-motile < 1) corre-
sponds to small and very quickly moving objects. This
upper-left part of the panel, close to the full ballistic
regime, is dominated by hydrodynamic screening and is
probably not relevant for marine particles, even for fast
sinking-fecal pellets, since their density differs from that
of seawater only by about 10%-20% [33, 34], see the red
Sinking speed U [μm/s]500635000(a)(d)103101102103101102500635000Sinking speed U [μm/s]motile, elongated, shear ON lb=2μm, Ub=50μm/s, α=3.3103101102103101102motile, elongated, shear OFF lb=2μm, Ub=50μm/s, α=3.310010-110-210-3Encounter efficiency η140°120°100°80°60°40°Mean colatitude ‹θ›103101102103101102non-motile, spherical lb=1μm, Ub=0, α=1(b)(e)(c)(f)Particle radius R [μm]Particle radius R [μm]Particle radius R [μm]16
this encounter rate is realized by hydrodynamic focusing,
which results in most bacteria attaching to the leeward
side (θ < 90°) of the sinking particle [compare Figs. 10(d)
and (e); see also Fig. A.3]. For small and slowly sinking
particles (R < 50 µm, U < 500 µm s−1), for which shear
dominates over rotational diffusion, more than 75% of
the interceptions occur on the leeward side of the par-
ticle. Furthermore, about 25% of the interceptions are
concentrated inside the 'Arctic circle' (θ < 23°), which
represents a more than five-fold increase as compared to a
uniform coverage of the particle. This leaves the southern
hemisphere depleted of bacteria, with almost no intercep-
tions below the 'tropic of Capricorn' (θ > 113°). Thus,
the leeward stagnation point is a flow-induced hotspot
where motile and elongated bacteria concentrate due to
shear. Since non-motile bacteria intercept the particles
on the upstream side [Fig. 10(f)], in the southern hemi-
sphere, we conclude that flow and shear lead to a bipolar
segregation of motile and non-motile marine bacteria on
the two sides of a sinking particle.
Although this work has assumed particles to be spher-
ical despite the variety of observed shapes exhibited by
marine snow aggregates [11], we expect the phenomena of
hydrodynamic focusing and screening of elongated bac-
teria to be robust to variation in shape. The focusing
and screening effects rely on different orientational re-
sponses of small rods upstream and downstream of the
particle - the key property of the flow that is required
for this fore -- aft symmetry breaking is the expansion of
the streamlines to the front of the particle and their re-
combination to the back, as well as the no-slip boundary
conditions on the particle surface. As long as such gen-
eral streamline organization is preserved, the effects here
described should be robust: while fluid parcels roll on
the particle surface (no slip), they stretch upstream of
the particle (streamline expansion) but compress down-
stream of the particle (streamline recombination). This
basic process will hold for objects at low Reynolds num-
bers with no-slip surfaces, and one would therefore expect
hydrodynamic focusing and screening of motile elongated
bacteria to occur for marine particles in general.
VII. CONCLUSIONS
In this work, we combined analytical and numerical
calculations to estimate the encounter rates between non-
motile and motile microorganisms of different morpholo-
gies and sinking particles in the ballistic regime relevant
for the most abundant small sinking particles. Previous
estimates have primarily focused on the diffusive regime,
effectively assuming that particles are much larger than
the bacterial run length. In the ballistic range, bacterial
reorientation becomes a significant factor influencing the
encounter process, while it is absent by necessity from
diffusive models. We have focused on the coupling be-
tween microorganism shape and fluid shear induced by
the particle, since shear is the dominant external factor
FIG. 11: Comparison of the encounter efficiencies between
motile elongated and non-motile spherical microorganisms (a)
and between motile elongated microorganisms with and with-
out shear (b) computed from the panels (a -- c) of Fig. 10. For
reference, we plot the absolute values of the sinking/raising
speeds as determined by the Stokes law (1) for sinking parti-
cles with densities ρp higher than the density of water ρw by
5 − 15% (in black font), such as the fast sinking fecal pellets,
as well as raising bubbles with ρp/ρw ≈ 0 (in white font). The
white dashed line denotes the ratio equal to unity.
and yellow lines in Fig. 11, which represent the Stokes
law (1). However, this hydrodynamic screening regime
may be relevant for interception by air bubbles, whose
vertical speed is high in view of their large density differ-
ence with seawater [35] (purple line in Fig. 11). Further-
more, comparing the shear on-off cases for motile elon-
gated bacteria [Fig. 11(b)], we find that the major im-
pact of shear is to reduce the encounter rates with small
particles sinking at intermediate or rapid rates by up to
a factor of 10. This reduction is a consequence of the
competition between the hydrodynamic screening of rods
upstream of the sinking particle and rotational diffusion.
It suggests that elongation-induced screening may be a
passive mechanism that allows motile elongated marine
microorganisms to prioritize slowly sinking aggregates,
at least in the quasi-ballistic particle size range.
For slowly sinking particles in the quasi-ballistic
regime (lower parts of the panels in Fig. 11), the observed
encounter efficiencies of motile marine bacteria are close
to the case without shear, consistent with earlier stud-
ies [12]. However, at the level of individual trajectories,
Particle radius R [μm]103101102(a)Sinking speed U [μm/s]500635000ratio ηmotile /ηnon-motileshear ON102101ρp/ρw=1.15ρp/ρw=1.10ρp/ρw=1.05ρp/ρw=0100ratio ηmotile /ηmotileshear ONshear OFF10-110-0.5Particle radius R [μm]103101102(b)Sinking speed U [μm/s]500635000100ρp/ρw=1.15ρp/ρw=1.10ρp/ρw=1.05ρp/ρw=0responsible for bacterial reorientation. We have shown
that the shape -- shear coupling can significantly affect
the encounter rate and attachment location on a particle
for both non-motile and motile microorganisms.
For non-motile organisms, shear from a sinking particle
can significantly alter the encounter rates of organisms
with different morphologies. For elongated organisms,
this influence occurs by aligning the cells' long axis tan-
gentially to the particle surface, and was experimentally
validated. When the timescale of rotational diffusion
is long with respect to particle interactions, shear from
sinking particles interacts with the aspect ratio of non-
motile organisms to potentially reduce encounter rates
by a factor proportional to the square of the aspect ratio.
As a result, encounters could exert evolutionary pressure
on non-motile cell morphology [36], favoring elongated
or disk-like shapes depending on whether encounters are
unfavorable or favorable, respectively.
For motile microorganisms, interactions with the shear
from a sinking particle give rise to two phenomena, hy-
drodynamic screening and focusing, that alter both the
rates and locations of encounters. Elongation helps or-
ganisms intercept slowly sinking particles but dramat-
ically reduces the encounter rate with rapidly sinking
particles. In contrast to rods, motile disks experience up-
stream focusing, leading to high efficiency at intercept-
ing rapidly sinking particles. From the perspective of
the particle, motile elongated microorganisms typically
attach to the leeward side of the particle, while motile
disks cover it more uniformly. Under realistic parameters
relevant to marine bacteria and sinking particles, which
include the effect of randomization of swimming direction
from rotational diffusion, hydrodynamic screening leads
to a ten-fold decrease in the interception rate of rapidly
sinking aggregates, as compared to motility without the
shear-induced reorientation. This reduction in encounter
rate suggests that elongation-induced screening may be a
passive mechanism that allows motile elongated marine
microorganisms to avoid rapidly moving particles. Last
but not least, motile elongated bacteria attach to the
leeward side of the particle, whereas non-motile bacteria
attach to the front. Thus, hydrodynamic focusing is a
physical source of heterogeneity in particle colonization
characterized by bipolar segregation of motile and non-
motile microorganisms, which may influence the degrada-
tion rate of marine snow aggregates. Whether in terms
of encounter rates or encounter locations, these results
indicate that the impact of shear reorientation cannot be
neglected when evaluating interactions between motile
organisms and sinking particles.
The dynamics of shear-driven reorientation are directly
relevant to the colonization of marine particles by bac-
teria. It is well established that motility can greatly en-
hance the encounter rate of bacteria with sinking parti-
cles [8, 24]. This enhancement is often estimated via the
ratio of the effective diffusivity due to motility and the
diffusivity due to Brownian motion, which can be as large
as 1000 for highly motile marine bacteria [8, 37, 38]. The
17
more accurate theory developed here, which accounts ex-
plicitly for the interaction between flow and motility in
elongated bacteria, refines this estimation in a manner
that depends on the particle size and sinking speed rela-
tive to the bacterial motility. For particles substantially
larger than the bacterial run length, the enhancement
in attachment due to motility estimated by the ratio of
effective diffusivities and neglecting the impact of shear
is increasingly more accurate. For the marine bacteria
modeled here, this corresponds to particles with radius
greater than approximately 1 mm. For smaller particles,
which form the bulk of particles in the ocean [39], this
work reveals that the enhancement in encounters result-
ing from motility is more moderate and is further reduced
as the particle sinking speed increases. In extreme cases,
potentially applicable to some bubbles, motility may con-
fer no benefit in encounter rates. However, in the context
of marine particles, motility still enhances encounters by
one to two orders of magnitude. Since the enhancement
in encounter rate due to motility is greater for slowly-
sinking particles, this also highlights the potential signif-
icance of neutrally buoyant particles [40] to motile bac-
teria. This fundamental knowledge of encounter rates
will be a valuable asset in future efforts to rationalize
the community composition on marine particles, and ul-
timately the role of different groups of bacteria in particle
degradation and the ocean's biological pump.
In a different domain, the mechanisms of hydrody-
namic focusing and screening of rods and disks here de-
scribed are relevant to the classical filtration problem [9],
because our results suggest that shear renders elongated
non-motile colloids more difficult to collect than oblate
ones. Furthermore, fabrication of Janus-type artificial
swimmers makes it possible to build microscale motile
objects with different shapes and swimming speed [32],
and these parameters could be tailored to enhance or sup-
press the focusing and screening effects. For example, the
efficiency in capturing moving spheres may be important
in applications such as targeted drug delivery [41] and
micromachine-enabled decontamination [42].
In summary, we have demonstrated that hydrody-
namic interactions between a small ellipsoid and a large
moving sphere break the fore -- aft symmetry of the flow
streamlines,
leading to practical consequences for mi-
croorganisms. This symmetry breaking is a conse-
quence of fluid expansion and recombination upstream
and downstream of the sphere, but is only revealed when
the full tensorial character of the velocity gradient is ac-
counted for, including its straining and rotational compo-
nents. Such asymmetric two-body couplings are ubiqui-
tous, since they arise when a small nonspherical particle
travels near a larger obstacle in a fluid; we have exper-
imentally verified their impact in the case of non-motile
elongated diatoms advected around an alginate bead. In
the context of swimming bacteria intercepting a sinking
particle, hydrodynamic focusing and screening have prac-
tical ecological impacts, but applications to other natural
or man-made systems are yet to be explored.
Appendix A: Supplementary Information
18
The Appendix is organized as follows: we provide linear stability analysis of the fixed points of the Jeffery equation
in Section A 1 and derive the limit cycle solutions and their period in Section A 2; these results were discussed
in Section III A of the Main Text. The velocity gradient for the Stokes flow is derived in Section A 3; its matrix
form was used in Eq. (13) in Section III B of the Main Text.
In Section A 4, we complement the discussion in
Section III B of the Main text by analyzing the structure of the velocity gradient on the stagnation lines and the sinking
particle surface. The subsequent Sections give details on the numerical simulations (Section A 5) and experimental
methods (Section A 6). Finally, the three additional figures supplement the Main Text as follows: Fig. A.1 shows
that the results presented in Fig. 5 of the Main Text are robust to variation in the cut-off threshold for rejecting the
out-of-plane components of rods in the simulations, Fig. A.2 quantifies the overestimate in the encounter efficiencies
as predicted by the classical diffusive arguments in the range of parameters discussed in Figs. 10 and 11 of the Main
Text, and Fig. A.3 provides an additional characterization of the landing distributions for the simulations presented
in Figs 10 of the Main Text.
1. Stability analysis of the fixed points of the Jeffrey equation
In this section, we analyze the linear stability of the fixed points of the Jeffrey Eq. (9)
p = (I − ppT)Aγp.
(A1)
This analysis will also yield the characteristic timescales of the convergence onto the asymptotically stable solutions.
As discussed in Section III A and in [18], the fixed points of Eq. (A1) are given by the real eigenvectors of Aγ. Let
λ be a normalized real eigenvector of Aγ with eigenvalue λ. Linearizing Eq. (A1) around λ by writing p = λ + ∆p,
where the perturbation ∆p lies in the tangent space to the sphere at λ, gives
∆p = −λ∆p + (I − λλT)(Aγ∆p).
(A2)
Eq. (A2) is a two-dimensional linear dynamical system whose stability can be classified using the standard trace-
determinant characterization. To be more explicit, we introduce the basis vectors {e1, e2} for the tangent space to
the sphere at λ
e1 = n × λ,
e2 = e1 × λ,
(A3)
where n is an arbitrary nonzero vector, noncolinear with λ. In that basis, the perturbation reads ∆p = α(t)e1 +β(t)e2
and the linearized system (A2) reduces to
(cid:35)(cid:20)α
(cid:21)
β
(cid:20)α
(cid:21)
β
(cid:35)
(cid:34) α
(cid:34)
(cid:20)α
(cid:21)
β
, = −λ
+
eT
1 Aγe1 eT
eT
2 Aγe1 eT
1 Aγe2
2 Aγe2
= Mλ
(cid:34)
1 Aγe1 − λ
eT
2 Aγe1
eT
(cid:35)
1 Aγe2
eT
2 Aγe2 − λ
eT
.
Mλ =
,
(A4)
(A5)
β
where
To evaluate the trace and determinant of Mλ, we first note that {e1, e2, λ} is an orthogonal basis for Aγ. In that
basis, Aγ takes the form
eT
Aγ =
1 Aγe2 0
1 Aγe1 eT
2 Aγe2 0
2 Aγe1 eT
eT
λTAγe1 λTAγe2 λ
,
(A6)
Let {λa, λb, λ} be the three eigenvalues of Aγ. The following identities follow from the above matrix representation
TrAγ = eT
1 Aγe1 + eT
det Aγ = λaλbλ = λ[(eT
2 Aγe2 + λ = λa + λb + λ = 0,
1 Aγe1)(eT
2 Aγe2) − (eT
1 Aγe2)(eT
2 Aγe1)].
(A7)
(A8)
where we used fluid incompressibility in the first equation. From the above, we derive the following formulae
eT
1 Aγe1 + eT
2 Aγe2) − (eT
1 Aγe2)(eT
2 Aγe2 = −λ,
2 Aγe1) = λaλb,
(eT
1 Aγe1)(eT
which imply the following expressions for the trace and determinant of M
From these expression, the eigenvalues of Mλ read
(TrMλ ±(cid:112)
λMλ± =
1
2
Since λ = −λa − λb this further simplifies to
TrMλ = −3λ,
det Mλ = 2λ2 + λaλb.
(−3λ ±(cid:112)
1
2
TrM 2 − 4 det Mλ) =
λ2 − 4λaλb).
(−3λ ±(cid:112)(λa − λb)2).
λMλ± =
1
2
19
(A9)
(A10)
(A11a)
(A11b)
(A12)
(A13)
We now use the eigenvalues (A13) to analyze the linear stability of the fixed points of the Jeffrey Eq. (A1). Let's
first consider the case when Aγ has all real eigenvalues λ1 < λ2 < λ3 with eigenvectors {λ1, λ2, λ3}, which are also
the fixed points of Eq. (A1). In this case, the eigenvalues of the linearized system (A4) are also real and Eq. (A13)
simplifies to
(A14)
Furthermore, since λ1 + λ2 + λ3 = 0 by the incompressibility, we must have λ1 = −λ1 < 0, λ3 > 0 and λ2 <
min(λ1, λ3). For the fixed point λ1, the eigenvalues are are always positive
λMλ± =
1
2
(−3λ ± λa − λb).
Mλ1± =
λ
1
2
(−3λ1 ± λ3 − λ2) =
[−3λ1 ± (λ3 − λ2)] =
1
2
1
2
[−3λ1 ± (−λ1 − 2λ2)] > 0,
implying that λ1 is a repulsive node. For the fixed point λ2, the eigenvalues are
Mλ2± =
λ
1
2
(−3λ2 ± λ1 − λ3) =
1
2
(−3λ2 ± λ2 + 2λ3) =
[−3λ2 ± (λ2 + 2λ3)].
1
2
Explicitly,
+ = λ3 − λ2 > 0,
Mλ2− = −2λ2 − λ3 = λ1 − λ2 < 0.
λ
Mλ2
λ
Thus, λ2 is a saddle point. Finally, for λ3, the eigenvalues are always negative
Mλ3± =
λ
1
2
(−3λ3 ± λ1 − λ2) =
1
2
[−3λ3 ± (λ3 + 2λ2)] < 0,
(A15)
(A16)
(A17a)
(A17b)
(A18)
implying that λ3 is an attracting node. We conclude that the asymptotically stable orientations of Eq. (A1) for the
case when Aγ has three real eigenvalues are given by ±λ3, since these are the only attracting fixed points on the
sphere of orientations. We can estimate the characteristic time τλ3 needed to converge onto the stable orientation λ3
as the inverse of the average of the eigenvalues λ
(A19)
1, λ3}.
We now consider the case when Aγ has a pair of complex conjugate eigenvalues and one real eigenvalue {λ1, λ∗
We write the complex eigenvalue as λ1 = λr
1. In this case, the only fixed point of the Jeffrey Eq. (A1) is given
by the real eigenvector λ3. We estimate the linear stability of this fixed point. The eigenvalues of the linearized
system (A4) become
1 + iλi
λ3.
λ3
Mλ3
+ + λ
Mλ3− )/2 =
3
2
Mλ3±
= −(λ
τ−1
[−3λ3 ±(cid:113)
Mλ3± =
λ
1
2
(λ1 − λ∗
1)2] = − 3
2
λ3 ± iλi
1.
(A20)
We see that, if the only real eigenvalue λ3 is positive, then λ3 is an attractive spiral. Otherwise, it's a repulsive spiral
and the asymptotic state of Eq. (A1) is given by a stable limit cycle, to be discussed in the next section in more
detail. The timescale associated with the convergence on or divergence away from λ3 is given by the absolute value
of the real part of λ
Mλ3±
τ−1
λ3
=
λ3.
3
2
(A21)
2. Limit cycle case
20
In the case when Aγ has complex eigenvalues λ1,2 and the real eigenvalue is negative λ < 0, the asymptotic solution
to the Jeffrey Eq. (A1) is given by a limit cycle. The limit cycle is the great circle perpendicular to the real eigenvector
p∗ of Aγ. To show this, we introduce the orthonormal basis
n1 = w × p∗/(cid:107)w × p∗(cid:107), n2 = n1 × p∗,
(A22)
where w is a random nonzero vector. Note that the two orthogonal vectors n1 and n2 span the plane of the great
circle perpendicular to the real eigenvector p∗. We look for solutions of the form
Plugging the above ansatz into the Jeffrey Eq. (A1) yields
p(t) = sin θ(t)n1 + cos θ(t)n2.
θ cos θn1 − θ sin θn2 = [I − (sin θn1 + cos θn2)(sin θnT
1 + cos θnT
2 )](sin θAγn1 + cos θAγn2).
To simplify the above expression, we introduce the following notation for the submatrix of Aγ
and project Eq. (A24) onto n1 and n2
M =
(cid:34)
(cid:35)
,
nT
nT
1 An1 nT
2 An1 nT
1 An2
2 An2
θ cos θ(t) = sin θM11 + cos θM12 − sin θ(sin2 θM11 + sin θ cos θM12 + cos θ sin θM21 + cos2 θM22),
θ sin θ(t) = − sin θM21 − cos θM22 + cos θ(sin2 θM11 + sin θ cos θM12 + cos θ sin θM21 + cos2 θM22).
(A26)
(A27)
We combine the two equations into a single one by taking a linear combination with weights cos θ and sin θ
θ cos2 θ(t) + θ sin2 θ(t) = θ = cos θ sin θM11 + cos2 θM12 − sin2 θM21 − sin θ cos θM22,
which further simplifies to
θ =
M11 − M22
2
M12 + M21
sin 2θ +
cos 2θ +
2
M12 − M21
.
2
Introducing A = M11 − M22, B = M12 + M21 and C = M12 − M21, we obtain
2 θ = A sin 2θ + B cos 2θ + C.
This is a first-order nonlinear differential equation. Since the nonlinear term is smooth, the unique (up to the 2π
period) solution exists, which validates the ansatz (A23) and proves the existence of a limit cycle.
We now explicitly calculate the period T of the limit cycle. To this end, integrate Eq. (A30) over T
(A23)
(A24)
(A25)
(A28)
(A29)
(A30)
(A31)
(A32)
(A33)
(A34)
(A35)
4π = CT +
(A sin 2θ + B cos 2θ)dt
= CT +
(A sin 2θ + B cos 2θ)
1
θ
dθ
= CT + 2
(1 −
C
A sin 2θ + B cos 2θ + C
)dθ.
0
0
(cid:90) T
(cid:90) 2π
(cid:90) 2π
(cid:90) 2π
0
We get the following equation for T
T =
This expression can be expressed as
2
dθ.
A sin 2θ + B cos 2θ + C
0
(cid:90) 2π
0
T = 2
1
A sin θ(cid:48) + B cos θ(cid:48) + C
dθ(cid:48) = 2
(cid:90) 2π
0
√
1
A2 + B2 sin θ + C
dθ,
where we changed variables twice using θ(cid:48) = 2θ and θ = θ(cid:48) +α, where sin α = B/
A2 + B2 and we used the periodicity
of the integrand to keep the integration limit as [0, 2π). The final integral can be evaluated using contour integration.
We first change the variables z =
A2 + B2eiθ, which yields
√
√
21
(cid:73)
T = 4
dz
z2 + 2iCz − (A2 + B2)
.
(A36)
We note that C 2 > A2 + B2 corresponds to A having complex eigenvalues, which is the case of interest. In this case,
A2 + B2 centered at the origin)
the integrand has one simple pole inside the integration contour (circle of radius
given by one of the roots of the integrand denominator. Applying the residue theorem, yields
√
√
T =
4π
C 2 − A2 − B2
.
This can be related to the original matrix A and its negative real eigenvalue λ as
(cid:112)2 det A/λ + λ2 − tr(A2)
4π
.
T =
Assuming the complex eigenvalues take the form λ1,2 = α ± iβ, this further simplifies to
T =
2π
β
.
(A37)
(A38)
(A39)
Therefore, the angular frequency of the limit cycle is given by the imaginary part of the complex eigenvalue. These
results agree with the analysis in [19] obtained using a different method.
3. Velocity gradient of the Stokes flow around a sphere
In this section, we compute the velocity gradient in Eq. (13) due to the Stokes flow around a sinking particle.
We first carry out the calculation in the curvilinear orthogonal coordinate basis {∂r, ∂θ, ∂φ} with the metric tensor
gij = diag(1, r2, r2 sin2 θ) and then transform to the usual orthonormal system {r, θ, φ}. The transformation between
the two systems is encoded in the Jacobian
J = diag(1, r, r sin θ).
In the curvilinear system {∂r, ∂θ, ∂φ}, the Stokes flow (2) reads
(cid:16)
(cid:17)
v = vr∂r + vθ∂θ = U cos θ
1 +
R3
2r3 − 3R
2r
∂r + U sin θ
(cid:16) − 1
r
+
R3
4r4 +
3R
4r2
(cid:17)
∂θ.
(A40)
(A41)
To compute the velocity gradient (1,1)-tensor Aij = ∇jvi, we note that the only nonzero Christoffel symbols are
φφ = −r sin2 θ,
θθ = −r, Γr
Γr
Γθ
rθ = Γθ
Γφ
rφ = Γφ
θr = 1/r, Γθ
φr = 1/r, Γφ
φφ = − sin θ cos θ,
θφ = Γφ
φθ = cot θ.
Using covariant differentiation, we find the velocity gradient tensor components (in the {∂r, ∂θ, ∂φ} basis)
∇rvr = ∂rvr, ∇rvθ = ∂rvθ + vθ/r, ∇rvφ = 0,
∇θvr = ∂θvr − rvθ, ∇θvθ = ∂θvθ + vr/r, ∇θvφ = 0,
∇φvr = 0, ∇φvθ = 0, ∇φvφ = vr/r + cot θvθ.
(A42)
(A43)
(A44)
(A45a)
(A45b)
(A45c)
Explicit calculation gives the following expressions for the tensor entries
∇rvr = U cos θ
∇rvθ = U sin θ
∇rvφ = 0,
∇θvr = −U sin θ
∇θvθ = U cos θ
∇θvφ = 0,
∇φvr = 0,
∇φvθ = 0,
∇φvφ = U cos θ
,
3R
2r2
(cid:17)
(cid:16) − 3R3
(cid:17)
(cid:16) 1
2r4 +
r2 − R3
r5 − 3R
(cid:16)
(cid:17)
(cid:16) − 1
2r3 − 3R
R3
4r4 +
1 +
2r3
R3
2r
+
r
3R
4r2
(cid:17)
+ U sin θ
r2 +
= U sin θ
(cid:16) − 1
(cid:16)
(cid:16) 1
r
(cid:17)
3R
4r3
R3
4r5 +
(cid:17)
2r4 − 3R
R3
2r2
(cid:17)
+ U sin θ
1 − R3
4r3 − 3R
4r
= U sin θ
+ U cos θ
+
= U cos θ
4r3
(cid:16) − 3R3
4r5 − 3R
(cid:17)
(cid:16) − 3R3
(cid:17)
(cid:16) 3R3
4r3 +
4r4 − 3R
3R
4r
4r2
,
,
(cid:16) 1
r
(cid:17)
+ U cos θ
(cid:16) − 1
r
(cid:17)
= U cos θ
(cid:16) 3R3
4r4 − 3R
4r2
(cid:17)
.
+
R3
4r4 +
3R
4r2
+
R3
2r4 − 3R
2r2
As a sanity check, we compute the flow divergence
which vanishes, as expected. In the matrix form, the above tensor reads (U = 1 and R = 1)
(cid:0) − 3
(cid:0) − 3
2r4 + 3
2r2
4r5 − 3
4r3
0
Aij = ∇jvi =
JAJ−1 =
3
4
(
1
r2 − 1
∇rvr + ∇θvθ + ∇φvφ = 0,
.
(cid:1) sin θ
(cid:1) cos θ
4r
(cid:1) cos θ (cid:0) − 3
(cid:1) sin θ
(cid:0) 3
4r3 + 3
4r4 − 3
r4 ) cos θ
4r2
0
0
tan θ
2
− r2+1
r2−1 tan θ −1
0
0
0 −1
,
0
0
(cid:0) 3
4r4 − 3
4r2
(cid:1) cos θ
Finally, we use the Jacobian J [Eq. (A40)] to express A in the orthonormal basis {r, θ, φ}
(cid:17)
,
22
(A46a)
(A46b)
(A46c)
(A46d)
(A46e)
(A46f)
(A46g)
(A46h)
(A46i)
(A47)
(A48)
(A49)
which yields Eq. (13), in agreement with the calculation in [21] where the velocity gradient was computed using a
different method.
4. Ellipsoids in the Stokes flow: stagnation lines and particle surface
The eigenvalues of the velocity gradient A [Eq. (13)] on the stagnation line (θ = 0, π) and the particle surface
(r = 1) have multiplicity greater than one. In this case, the analysis of Section III A does not directly apply, yet these
special locations will be important for the encounter process of non-motile microorganisms, which can only approach
the sinking particle near the stagnation line θ = π. On the stagnation lines θ = 0, π, Eq. (13) reduces to
Aij(r, θ = 0, π, φ) = ± 3
4
(cid:16) 1
r2 − 1
r4
(cid:17)2 0
0
0 −1 0
0 0 −1
,
(A50)
where +/− corresponds to θ = 0 and θ = π, respectively. This simple diagonal structure implies that on the upstream
stagnation line (θ = π) rods align tangentially to the sinking particle, while on the downstream stagnation line (θ = 0),
rods align vertically. This picture can be inferred from Fig. 3(b) by taking the limit ρ → 0. Since Aγ=−1 = −AT,
we immediately obtain the response of disks. Disks align tangentially to the particle surface for θ = π (with axis
of symmetry in the vertical direction), while they lie in the r − θ plane for θ = 0. Therefore, non-motile rods or
disks approaching the sinking particle along the θ = π stagnation line orient with their longer dimension tangential
to the particle surface. We now compute A on the particle surface to see if shear tends to maintain such a tangential
orientation. At r = 1, the only nonzero component of the velocity gradient is Aθr(r = 1, θ, φ) = − 3
2 sin θ. This
structure implies that the tangential orientations of rods (symmetry axis along θ − φ) and disks (symmetry axis along
r) are the null vectors. Thus, to zeroth order, shear maintains the tangential orientation of rods and disks as they
are advected around the sinking particle.
23
FIG. A.1: Additional comparison between experiments with non-motile elongated diatoms (a) and simulations (b -- d) shown in
Fig. 5 of the Main Text. In simulations, we reject rods with out-of-plane components larger than: 15° (b), 30° (c) and 45° (d).
Panels (a) and (c) are the same as panels (b,c) in Fig. 5 of the Main Text.
FIG. A.2: Ratios of the encounter efficiencies with and without the impact of shear, ηmotile
shear OFF (b), and
the encounter efficiency based on the classical diffusive calculation ηmotile
diffusive; all parameters are the same as in Fig. 10 and
diffusive = 4Sh/Pe, where
Fig. 11 of the Main Text.
Sh and Pe are the Sherwood and P´eclet number, respectively [25]. For the Sherwood number, we used the following formula
valid for low Reynolds number Sh = 0.5[1 + (1 + 2Pe)1/3]. For the P´eclet number, we took Pe = U R/Db, with the bacterial
; for the parameters used, Db = 5 × 10−5cm2/s. As discussed in Section VI, the
diffusivity Db = 0.5U 2
diffusive encounter efficiency overestimates the ballistic one and only for the largest sinking particles considered here the two
descriptions start to become comparable.
In the presence of flow, the diffusive encounter efficiency is given by ηmotile
shear ON (a) and ηmotile
b τd, where τd = D−1
r
5. Methods: numerical simulations
Time stepping. To numerically integrate the ballistic model [Eq. (3)], we discretized the equations of motion
using the classical Runge -- Kutta method (RK4). Depending on the sinking speed, the time-step was chosen between
∆t = 0.075τb for U ∼ Ub and ∆t = 0.005τb for U ∼ 100Ub, where τb = lb/Ub is the time needed for the bacterium
to travel distance equal to its bodylength. To integrate the quasi-ballistic model with rotational diffusion [Eq. (20)],
we used the stochastic version of the Euler method; at each time step, the diffusive term is discretized by sampling
a 3 × 1 vector with normally distributed entries with zero mean and variance 2Dr∆t. With the rotational diffusion
coefficient Dr = 0.25s−1, the time step varied between 0.04 − 0.2ms.
Estimation of the encounter efficiency. For a given sinking particle size R and sinking speed U , we estimated
the encounter efficiency by discretizing Eq. (7). We typically sampled the encounter probability P (ρ) on a non-uniform
grid to resolve the accumulation of P (ρ) near the accessibility region for slowly sinking particles or near the centerline
for fast sinking particles (see Figs. 7 and 8); the number of points on the ρ-grid was always at least 50 for motile
bacteria and 10 for non-motile bacteria. Once the estimate of P (ρ) had been obtained, the integral in Eq. (7) was
evaluated using the trapezoidal rule.
To estimate the encounter probability P (ρ) starting in the initial plane z = −6R at distance ρ away from the
centerline for random initial orientations, we considered an ensemble of initial orientations by sampling along a
spherical spiral. Such a sampling gives an approximately uniform distribution of points on a unit sphere of orientations.
By using spherical spiral to sample initial orientations rather than choosing them randomly (that is, choosing 3 × 1
vectors with normal entries with zero mean and normalizing them to unit length), we obtained faster convergence
(a)experiment(c)theory, 30° cutoff 100.5Sine of the angle with verticalx/R0-22-11z/R210-2-1(d)(b)theory, 15° cutoff theory, 45° cutoff x/R0-22-11x/R0-22-11x/R0-22-11(a)103101102Sinking speed U [μm/s]500635000ratio ηmotile /ηmotileshear OFFdiffusive103101102(b)ratio ηmotile /ηmotileshear ONdiffusive10010-110-210-3Particle radius R [μm]Particle radius R [μm]24
Formally, θf is defined as the integral 2π(cid:82) θf
FIG. A.3: Additional characterization of the landing distribution function ξ(θ) for the simulations presented in Fig. 10 of the
Main Text. Here, we look at the colatitude θf such that the fraction f of the interception positions lies in between 0° < θ < θf .
ξ(θ) sin θdθ = f . We display the results for three fractions, f = 0.25 (top row),
f = 0.5 (middle row) and f = 0.75 (bottom row). For reference, we note that a uniform coverage of the sphere corresponds to
θ0.25 = 60°, θ0.5 = 90° and θ0.75 = 120°.
0
by avoiding random clustering of points on the unit sphere. The number of initial orientations was chosen high
enough to guarantee that the solid angle the sinking particle extended at the initial bacterial location contained at
least five initial orientations (the number of blue dots in the inset in Fig. 1 was always at least five). For such an
angular resolution, the number of initial orientations varied between O(102) for fast sinking particles up to O(104)
for slowly sinking particles, for which the accessibility region with P (ρ) > 0 was largest. In general, to estimate the
encounter efficiency η for a given (R, U ) pair, we simulated about O(104) trajectories for fast sinking particles and up
to O(106) trajectories for slowly sinking particles. In total, due to scanning the (R, U ) parameter space in different
shear ON/OFF configurations, the this work summarizes the results of simulating about O(108) bacterial trajectories.
Interception criterion. The sinking particle was assumed to be a perfect absorber: geometric overlap between
any part of a bacterium and the particle was counted as an encounter. In simulations, for simplicity, we computed this
geometric overlap by approximating elongated bacteria (α > 1) of length lb by a cylinder with spherical caps. The
cylinder length, including caps, is lb and its width is lb/α. With this simplification, determining the interception is
equivalent to determining the distance between the cylinder centerline and the particle center. Similarly, the geometry
of oblate particles (α < 1) was approximated by considering four cylinders (with spherical caps) of length lb and width
lbα. The centerlines of the cylinders lie in a plane, the centerline midpoints coincide and the centerlines are rotated at
angle 45° - the four cylinders form two crosses rotated by 45°. With this simplification, determining the interception
is equivalent to determining the distance between the four cylinder centerlines and the particle center.
Estimation of the distribution of interception locations. To estimate ξ(θ) for a given (R, U ) pair, we
(a)500635000Sinking speed U [μm/s]motile, elongated, shear ON lb=2μm, Ub=50μm/s, α=3.3motile, elongated, shear OFF lb=2μm, Ub=50μm/s, α=3.3103101102103101102103101102non-motile, spherical lb=1μm, Ub=0, α=1(b)(c)180°150°120°90°30°0°Colatitude of 25% coverage60°Particle radius R [μm]103101102103101102103101102103101102103101102103101102500635000Sinking speed U [μm/s]500635000Sinking speed U [μm/s]180°150°120°90°30°0°60°180°150°120°90°30°0°60°Colatitude of 50% coverageColatitude of 75% coverageParticle radius R [μm]Particle radius R [μm]considered the ensemble of the endpoints of trajectories that resulted in the interception. As described above, this
ensemble resulted from scanning the ρ-range, the initial position at distance ρ away from the particle centerline in the
initial z = −6R-plane, as well as uniform initial orientations. This ensemble yielded a histogram of the interception
colatitudes θ. During construction of this histogram, the counts for each scanned position ρ were further weighted by
ρ and the ρ-grid spacing, to account for the number of initial positions at distance ρ being proportional to ρ (circles
of radius ρ) as well as the non-uniformity of the ρ-grid. Such prepared histogram of θ-counts (30 bins, bin width 6°),
normalized to a probability density function over a unit sphere, was used as the estimate of ξ(θ).
25
6. Methods: experiments
Cell cultured. Phaeodactylum tricornutum cells (strain CCMP2561) were cultured in f/2 medium (Guillard and
Ryther 1962) mixed with artificial seawater. Artificial seawater was prepared by dissolving 35 g of artificial sea salt
(Instant Ocean, Spectrum Brands) in 1 L double distilled water, filtered through a 0.2 µm filter and autoclaved.
Cultures were propagated in 18° C in AlgaeTron AG 230 PSI (Photon Systems Instruments) with 14h/10h light/dark
cycle. For the experiments, cells in the exponential growing phase were used. Cell length and width were 21.2±2.4 µm
(n = 14) and 3.12 ± 0.57 µm (n = 14), respectively, as measured by phase microscopy.
Experimental procedure. Alginate beads were prepared using a mix of sodium alginate salt from brown algae
(1.5% w/v, medium viscosity; Sigma) with 50 mM ethylenediamine tetra acetic acid (EDTA) in double distilled water
(DDW). Beads were prepared by dripping the alginate solution from a 1 mL syringe at rate of 60 µL min−1 from a
height of 20 cm to beaker containing 0.5 M CaCl2 in DDW. The CaCl2 solution was stirred at 300 rpm, using a
magnetic stir-bar. Flow dynamics were studied in microfluidic chip (Sticky-Slide 0.4 IBIDI). For the experiment,
single bead was trapped at the center of the channel using a glass cover slide. A syringe pump (Harvard PHD2000)
was then used to feed the channels with artificial sea water at the desired flow speed (160 µm s−1). Chanel was
visualized using a Nikon (eclipse TI-2) microscope at magnification of 4 × 1.5 × 10 (60×) and 20 fps using Orca flash
4.0 (Hamamatsu) camera.
Image analysis was performed on 50 consecutive images using ImageJ (Rueden, C. T et al.
(2017)). In general 1098 ± 50 cells/image were measured. To extract the orientation from each cell, median image
intensity was calculated using 'Stacks/Z Projection' function and subtract from all images.
'FFT/Bang pass' filter
was used with small cutoff of 1 pxl and big cutoff of 10 pxl.
'Minimum' filter was used with 1 pxl cutoff. Data
was transformed to binary using 'Make binary' function with default parameters. Finally, 'Analyze particle' function
(Particle 30-1000 pxl) was used to collect the orientation data.
Image analysis.
[1] F. Lundell, L. D. Soderberg, and P. H. Alfredsson. Fluid mechanics of papermaking. Annu. Rev. Fluid Mech., 43(1):195 --
217, 2011.
[2] G. Falkovich, A. Fouxon, and M. G. Stepanov. Acceleration of rain initiation by cloud turbulence. Nature, 419(6903):151 --
154, 2002.
[3] G. A. Jackson, A. M. Waite, and P. W. Boyd. Role of algal aggregation in vertical carbon export during SOIREE and in
other low biomass environments. Geophys. Res. Lett., 32(13):L13607, 2005.
[4] J. S. Guasto, R. Rusconi, and R. Stocker. Fluid mechanics of planktonic microorganisms. Annu. Rev. Fluid Mech.,
44(1):373 -- 400, 2012.
[5] A. R. Longhurst and W. G. Harrison. The biological pump: profiles of plankton production and consumption in the upper
ocean. Prog. Oceanogr., 22(1):47 -- 123, 1989.
[6] H. W. Ducklow, D. K. Steinberg, and K. O. Buesseler. Upper ocean carbon export and the biological pump. Oceanography,
14(4):50 -- 58, 2001.
[7] L. Karp-Boss, E. Boss, and P. A. Jumars. Nutrient fluxes to planktonic osmotrophs in the presence of fluid motion.
Oceanogr. Mar. Biol. Annu. Rev., 34:71 -- 107, 1996.
[8] T. Kiørboe, H.-P. Grossart, H. Ploug, and K. Tang. Mechanisms and rates of bacterial colonization of sinking aggregates.
Appl. Environ. Microbiol., 68(8):3996 -- 4006, 2002.
[9] S. K. Friedlander. Mass and heat transfer to single spheres and cylinders at low Reynolds numbers. AIChE J., 3(1):43 -- 48,
1957.
[10] K. Son, F. Menolascina, and R. Stocker. Speed-dependent chemotactic precision in marine bacteria. Proc. Natl. Acad. Sci.
U.S.A., 113(31):8624 -- 8629, 2016.
[11] A. B. Bochdansky, M. A. Clouse, and G. J. Herndl. Dragon kings of the deep sea: marine particles deviate markedly from
the common number-size spectrum. Sci. Rep., 6:22633, 2016.
[12] T. Kiørboe and G. A. Jackson. Marine snow, organic solute plumes, and optimal chemosensory behavior of bacteria.
Limnol. Oceanogr., 46(6):1309 -- 1318, 2001.
26
[13] Marcos, H. C. Fu, T. R. Powers, and R. Stocker. Separation of microscale chiral objects by shear flow. Phys. Rev. Lett.,
102(15):158103, 2009.
[14] R. Rusconi, J. S. Guasto, and R. Stocker. Bacterial transport suppressed by fluid shear. Nat. Phys., 10(3):212 -- 217, 2014.
[15] M. T. Barry, R. Rusconi, J. S. Guasto, and R. Stocker. Shear-induced orientational dynamics and spatial heterogeneity in
suspensions of motile phytoplankton. J. Royal Soc. Interface, 12(112):20150791, 2015.
[16] A. L. Alldredge and C. C. Gotschalk. Direct observations of the mass flocculation of diatom blooms: characteristics,
settling velocities and formation of diatom aggregates. Deep Sea Res. A, 36(2):159 -- 171, 1989.
[17] G. B. Jeffery. The motion of ellipsoidal particles immersed in a viscous fluid. Proc. R. Soc. Lond. A, 102(715):161 -- 179,
1922.
[18] M. Junk and R. Illner. A new derivation of Jeffery's equation. J. Math. Fluid Mech., 9(4):455 -- 488, 2007.
[19] F. P. Bretherton. The motion of rigid particles in a shear flow at low Reynolds number. J. Fluid Mech., 14(2):284 -- 304,
1962.
[20] G. K. Batchelor. Mass transfer from a particle suspended in fluid with a steady linear ambient velocity distribution. J.
Fluid Mech., 95(2):369 -- 400, 1979.
[21] T. Kiørboe and A. W. Visser. Predator and prey perception in copepods due to hydromechanical signals. Mar. Ecol. Prog.
Ser., 179:81 -- 95, 1999.
[22] S. Hess, L. Eme, A. J. Roger, and A. G. B. Simpson. A natural toroidal microswimmer with a rotary eukaryotic flagellum.
Nat. microbiol., 2019.
[23] J.-B. Raina, V. I. Fernandez, B. Lambert, R. Stocker, and J. R. Seymour. The role of microbial motility and chemotaxis
in symbiosis. Nat. Rev. Microbiol., 17:284 -- 294, 2019.
[24] A. W. Visser and T. Kiørboe. Plankton motility patterns and encounter rates. Oecologia, 148(3):538 -- 546, 2006.
[25] T. Kiørboe. A mechanistic approach to plankton ecology. Princeton University Press, 2008.
[26] S. Chandrasekhar. Stochastic problems in physics and astronomy. Rev. Mod. Phys., 15(1):1 -- 89, 1943.
[27] D. Saintillan and M. J. Shelley.
Instabilities and pattern formation in active particle suspensions: kinetic theory and
continuum simulations. Phys. Rev. Lett., 100(17):178103, 2008.
[28] R. N. Bearon and A. L. Hazel. The trapping in high-shear regions of slender bacteria undergoing chemotaxis in a channel.
J. Fluid Mech., 771:R3, 2015.
[29] H. C. Berg and E. M. Purcell. Physics of chemoreception. Biophys. J., 20(2):193 -- 219, 1977.
[30] E. Secchi, A. Vitale, G. L. Mino, V. Kantsler, L. Eberl, R. Rusconi, and R. Stocker. The effect of flow on swimming
bacteria controls the initial colonization of curved surfaces (submitted). 2019.
[31] G. L. Mino, M. Baabour, R. Chertcoff, G. Gutkind, E. Cl´ement, H. Auradou, and I. Ippolito. E. coli accumulation behind
an obstacle. Adv. Microbiol., 8:451 -- 464, 2018.
[32] C. Bechinger, R. Di Leonardo, H. Lowen, C. Reichhardt, G. Volpe, and G. Volpe. Active particles in complex and crowded
environments. Rev. Mod. Phys., 88(4):045006, 2016.
[33] K. W. Bruland and M. W. Silver. Sinking rates of fecal pellets from gelatinous zooplankton (Salps, Pteropods, Doliolids).
Mar. Biol., 63(3):295 -- 300, 1981.
[34] P. D. Komar, A. P. Morse, L. F. Small, and S. W. Fowler. An analysis of sinking rates of natural copepod and euphausiid
fecal pellets. Limnol. Oceanogr., 26(1):172 -- 180, 1981.
[35] M. E. Weber, D. C. Blanchard, and L. D. Syzdek. The mechanism of scavenging of waterborne bacteria by a rising bubble.
Limnol. Oceanogr., 28(1):101 -- 105, 1983.
[36] A. Persat, H. A. Stone, and Z. Gitai. The curved shape of Caulobacter crescentus enhances surface colonization in flow.
Nat. Commun., 5:3824, 2014.
[37] B. S. Lambert, V. I. Fernandez, and R. Stocker. Motility drives bacterial encounter with particles responsible for carbon
export throughout the ocean. Limnol. Oceanogr. Lett., in press, 2019.
[38] H. C. Berg. Random walks in biology. Princeton University Press, 1993.
[39] L. Stemmann and E. Boss. Plankton and particle size and packaging: From determining optical properties to driving the
biological pump. Ann. Rev. Mar. Sci., 4(1):263 -- 290, 2012.
[40] X. Mari, U. Passow, C. Migon, A. B. Burd, and L. Legendre. Transparent exopolymer particles: Effects on carbon cycling
in the ocean. Prog. Oceanogr., 151:13 -- 37, 2017.
[41] J. Wang and W. Gao. Nano/microscale motors: biomedical opportunities and challenges. ACS Nano, 6(7):5745 -- 5751,
2012.
[42] W. Gao and J. Wang. The environmental impact of micro/nanomachines: a review. ACS Nano, 8(4):3170 -- 3180, 2014.
|
1810.07623 | 1 | 1810 | 2018-10-17T15:38:01 | Searching for collective behavior in a small brain | [
"physics.bio-ph",
"cond-mat.dis-nn",
"q-bio.NC"
] | In large neuronal networks, it is believed that functions emerge through the collective behavior of many interconnected neurons. Recently, the development of experimental techniques that allow simultaneous recording of calcium concentration from a large fraction of all neurons in Caenorhabditis elegans - a nematode with 302 neurons - creates the opportunity to ask if such emergence is universal, reaching down to even the smallest brains. Here, we measure the activity of 50+ neurons in C. elegans, and analyze the data by building the maximum entropy model that matches the mean activity and pairwise correlations among these neurons. To capture the graded nature of the cells' responses, we assign each cell multiple states. These models, which are equivalent to a family of Potts glasses, successfully predict higher statistical structure in the network. In addition, these models exhibit signatures of collective behavior: the state of single cells can be predicted from the state of the rest of the network; the network, despite being sparse in a way similar to the structural connectome, distributes its response globally when locally perturbed; the distribution over network states has multiple local maxima, as in models for memory; and the parameters that describe the real network are close to a critical surface in this family of models. | physics.bio-ph | physics | Searching for collective behavior in a small brain
Xiaowen Chen,1 Francesco Randi,1 Andrew M. Leifer,1,2 and William Bialek1,3,4
1Joseph Henry Laboratories of Physics, 2Princeton Neuroscience Institute,
and 3Lewis-Sigler Institute for Integrative Genomics, Princeton University, Princeton, NJ 08544
4Initiative for the Theoretical Sciences, The Graduate Center,
City University of New York, 365 Fifth Ave., New York, NY 10016
(Dated: October 18, 2018)
In large neuronal networks, it is believed that functions emerge through the collective behavior
of many interconnected neurons. Recently, the development of experimental techniques that allow
simultaneous recording of calcium concentration from a large fraction of all neurons in Caenorhab-
ditis elegans -- a nematode with 302 neurons -- creates the opportunity to ask if such emergence is
universal, reaching down to even the smallest brains. Here, we measure the activity of 50+ neurons
in C. elegans, and analyze the data by building the maximum entropy model that matches the
mean activity and pairwise correlations among these neurons. To capture the graded nature of the
cells' responses, we assign each cell multiple states. These models, which are equivalent to a family
of Potts glasses, successfully predict higher statistical structure in the network. In addition, these
models exhibit signatures of collective behavior: the state of single cells can be predicted from the
state of the rest of the network; the network, despite being sparse in a way similar to the structural
connectome, distributes its response globally when locally perturbed; the distribution over network
states has multiple local maxima, as in models for memory; and the parameters that describe the
real network are close to a critical surface in this family of models.
I.
INTRODUCTION
The ability of the brain to generate coherent thoughts,
percepts, memories, and actions depends on the coordi-
nated activity of large numbers of interacting neurons. It
is an old idea in the physics community that these collec-
tive behaviors in neural networks should be describable
in the language of statistical mechanics [1 -- 3]. For many
years it was very difficult to connect these ideas with
experiment, but new opportunities are offered by the re-
cent emergence of methods to record, simultaneously, the
electrical activity of large numbers of neurons [4 -- 9]. In
particular, it has been suggested that maximum entropy
models [10] provide a path to construct a statistical me-
chanics description of network activity directly from real
data [11], and this approach has been pursued in the
analysis of the vertebrate retina as it responds to natural
movies and other light conditions [11 -- 14], the dynamics
of the hippocampus during exploration of real and vir-
tual environments [15 -- 17], and the coding mechanism of
spontaneous spikes in cortical networks [18 -- 20].
Maximum entropy models that match low order fea-
tures of the data, such as the mean activity of individual
neurons and the correlations between pairs, make quan-
titative predictions about higher order structures in the
network, and in some cases these are in surprisingly de-
tailed agreement with experiment [14, 17]. These models
also illustrate the collective character of network activ-
ity. In particular, the state of individual neurons often
can be predicted with high accuracy from the state of the
other neurons in the network, and the models that are in-
ferred from the data are close to critical surfaces in their
parameter space, which connects with other ideas about
the possible criticality of biological networks [21 -- 23].
nomena in networks of neurons has been focused on verte-
brate brain, with neurons that generate discrete, stereo-
typed action potentials or spikes [24]. This discreteness
suggests a natural mapping into an Ising model, which is
at the start of the maximum entropy analyses, although
one could imagine alternative approaches. What is not
at all clear is whether these approaches could capture
the dynamics of networks in which the neurons generate
graded electrical responses. An important example of
this question is provided by the nematode Caenorhabdi-
tis elegans, which does not have the molecular machinery
needed to generate conventional action potentials [25].
The nervous system of C. elegans has just 302 neu-
rons, yet the worm can still exhibit complex neuronal
functions: locomotion, sensing, nonassociative and asso-
ciative learning, and sleep-wake cycles [26 -- 29]. All of
the neurons are "identified," meaning that we can find
the cell with a particular label in every organism of the
species, and in some cases we can find analogous cells
in nearby species [30]. In addition, this is the only or-
ganism in which we know the entire pattern of connec-
tions among the cells, usually known as the (structural)
connectome [31]. The small size of this nervous system,
together with its known connectivity, has always made
it a tempting target for theorizing, but relatively little
was known about the patterns of electrical activity in
the system. This has changed dramatically with the de-
velopment of genetically encodable indicator molecules,
whose fluorescence is modulated by changes in calcium
concentration, a signal which in turn follows electrical
activity [32]. Combining these tools with high resolution
tracking microscopy opens the possibility of recording the
activity in the entire C. elegans nervous system as the an-
imal behaves freely [7 -- 9].
Thus far, almost all discussion about collective phe-
In this paper we make a first try at the analysis of
8
1
0
2
t
c
O
7
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
3
2
6
7
0
.
0
1
8
1
:
v
i
X
r
a
2
FIG. 1: Schematics of data acquisition and processing. (a) Examples of the raw images acquired through the 10× (scale bar
equals 100µm) and 40× (scale bar equals 10µm) objectives. The body of the nematode is outlined with light green curves.
(b) The intensity of the nuclei-localized fluorescent protein tags -- the calcium-sensitive GCaMP and the control fluorophore
RFP -- are measured as functions of time. Photobleaching occurs on a longer time scale than the intracellular calcium dynamics,
which allows us to perform photobleaching correction by dividing the raw signal with its exponential fit, resulting in the signals
of panel (c). (d) The normalized ratio of the photobleaching-corrected intensity, f , is a proxy for the calcium concentration in
each neuron nuclei (dark grey). As described in the text, this signal is discretized using the denoised time derivative f ; we use
three states, marked as red, blue, and black after smoothing (lightly offset for ease of visualization). (e) The time derivative f ,
extracted using total-variation regularized differentiation.
experiments in C. elegans using the maximum entropy
methods that have been so successful in other contexts.
Experiments are evolving constantly, and in particular we
expect that recording times will increase significantly in
the near future. To give ourselves the best chance of say-
ing something meaningful, we focus on sub -- populations
of up to fifty neurons, in immobilized worms where sig-
nals are most reliable. We find that, while details dif-
fer, the same sorts of models, which match mean activity
and pairwise correlations, are successful in describing this
very different network. In particular, the models that we
learn from the data share topological similarity with the
known structural connectome, allow us to predict the ac-
tivity of individual cells from the state of the rest of the
network, and seem to be near a critical surface in their
parameter space.
II. DATA ACQUISITION AND PROCESSING
Following methods described previously [7, 8], nema-
todes Caenorhabditis elegans were genetically engineered
to expressed two fluorescent proteins in all of their neu-
rons, with tags that cause them to be localized to the
nuclei of these cells. One of these proteins, GCaMP6s,
fluoresces in the green with an intensity that depends
on the surrounding calcium concentration, which follows
the electrical activity of the cell and in many cases is
the proximal signal for transmission across the synapses
to other cells [32]. The second protein, RFP, fluoresces
in the red and serves as a position indicator of the nu-
clei as well as a control for changes in the visibility of
the nuclei during the course of the experiment. Parallel
control experiments were done on worms engineered to
express GFP and RFP, neither of which should be sen-
sitive to electrical activity. Although our ultimate goal
is to understand neural dynamics in the freely moving
animal, as a first step we study worms that are immobi-
lized with polystyrene beads, to reduce motion-induced
artifacts [33].
As described in Ref. [7], the fluorescence is excited us-
ing lasers. A spinning disk confocal microscope and a
high-speed, high-sensitivity Scientific CMOS (sCMOS)
camera records red- and green-channel fluorescent image
of the head of the worm at a rate of 6 brain-volumes
per second at a magnification of 40×; a second imag-
ing path records the position and posture of the worm
at a magnification of 10×, which are used in the track-
ing of the neurons across different time frames. The raw
data thus are essentially movies, and by using a custom
machine-learning approach -- Neuron Registration Vector
Encoding [8] -- we are able to reduce the data to the green
and red intensities for each neuron i, I g
i (t) and I r
i (t).
As indicated in Fig. 1b, the fluorescence intensity un-
(a) (c)0100200300400500time (s)020040060080010001200fluorescence intensity per area (a.u.)(d)f0100200300400500time (s)0.511.5I0100200300400500time (s)0.511.50100200300400500time (s)-0.0500.05(e)f.(b)10× IR40× GCaMP40× RFP10× RFP3
i (t) and ¯I r
divide these out to recover normalized intensities in each
channel and each cell, ¯I g
i (t). Finally, to reduce
instrumental and/or motion induced artifacts, we con-
sider the ratio of the normalized intensities as the signal
for each neuron, i.e. fi(t) = ¯I g
In
this normalization scheme, if the calcium concentration
remains constant, then fi(t) = 1.
i (t) (Fig. 1d).
i (t)/ ¯I r
Our goal is to write a model for the joint probability
distribution of activity in all of the cells in the network.
To stay as close as possible to previous work, at least in
this first try, it makes sense to quantize the activity into
discrete states. One possibility is to discretize based on
the magnitude of the fluorescence ratio fi(t). But this
is problematic, since even in "control" worms where the
fluorescence signal should not reflect electrical activity,
variations in different cells are correlated; this is illus-
trated in Fig. 2a, where we see that the distribution of
mutual information between fi(t) and fj(t), across all
pairs (i, j), is almost the same in control and experimen-
tal worms. A closer look at the raw signal suggests that
normalizing by the RFP intensity is not enough to correct
for occasional wobbles of the worm; this causes the dis-
tribution of the fluorescence ratio to be non-stationary,
and generates spurious correlations. This suggests that
(instantaneous) fluorescence signals are not especially re-
liable, at least given the current processing methods and
the state of our experiments. An alternative is to look at
the derivatives of these signals, which by definition suffer
from the global noise only at a few instances; now there
is very little mutual information between fi(t) and fj(t)
in the control worms, and certainly much less than in the
experimental worms, as seen in Fig. 2b.
To give ourselves a bit more help in isolating a mean-
ingful signal, we denoise the time derivatives. The op-
timal Bayesian reconstruction of the underlying time
derivative signal u(t) combines a description of noise in
the raw fluorescence signal f (t) with some prior expec-
tations about the signal u itself. We approximate the
noise in f as Gaussian and white, which is consistent
with what we see at high frequencies, and we assume
that the temporal variations in the derivative are expo-
nentially distributed and only weakly correlated in time.
Then maximum likelihood reconstruction is equivalent to
minimizing
(cid:90) T
0
dtAu − f2 ,
(2)
(cid:90) T
0
F (u) =
τf
σf
dt u +
1
2σ2
nτn
where A is the antiderivative operator, the combination
σ2
nτn is the spectral density of noise floor that we see in f
at high frequencies, while σf is the total standard devia-
tion of the signal and τf is the typical time scale of these
variations; for more on these reconstruction methods see
Ref. [35]. We determine the one unknown parameter τf
by asking that, after smoothing, the cumulative power
spectrum of the residue Au − f has the least root mean
square difference from the cumulative power spectrum of
the extrapolated white noise.
As an example, Fig. 1e shows the smooth derivative
FIG. 2: Comparison of pairwise mutual information distribu-
tion for the calcium-sensitive GCaMP worms and the GFP
control worms. Mutual information is estimated using bin-
ning and finite-sample extrapolation methods as described
in [34] for all pairs of neurons. For the normalized fluores-
cence ratio, f , the distribution of the mutual information,
P (I(fi; fj)), exhibits little difference between the calcium-
sensitive GCaMP worm and the GFP control worm (panel
(a)). In comparison, for the time derivative of the normalized
f , the distribution of the mutual informa-
fluorescence ratio,
tion, P (I( fi; fj)), is peaked around zero for the GFP control
worm, while the distribution is wide for the calcium-sensitive
GCaMP worm (panel (b)). This observation suggests that
fi, is more informative
time derivative of fluorescence ratio,
than its magnitude, fi.
dergoes photobleaching, fortunately on much longer time
scale than the calcium dynamics. Thus, we can extract
the photobleaching effect by modeling the observed fluo-
rescence intensity with an exponential decay:
Ig(t) = Sg(t)(1 + ηg)(e−t/τg + Ag)
Ir(t) = Sr(t)(1 + ηr)(e−t/τr + Ar)
(1)
Here, Sg(t) and Sr(t) are the true signals corresponding
to the calcium concentration, ηg and ηr are stochastic
variables representing the noise due to the laser and the
camera, τg and τr are the characteristic time for photo-
bleaching of the two fluorophores, and Ag and Ar repre-
sent nonnegative offsets due to a population of unbleach-
able fluorophores, or regeneration of fluorescent states
under continuous illumination.1
For each neuron, we fit the observed fluorescence in-
tensities to Eqs (1) with Sg(t) = S0
g and ηg = 0, and sim-
ilarly for Sr(t). As shown by the black lines in Fig. 1b,
this captures the slow photobleaching dynamics; we then
1 One may worry that a constant "background" fluorescence
should be subtracted from the raw signal, rather than contribut-
ing to a divisive normalization. In our data, this background sub-
traction leads to strongly non-stationary noise in the normalized
intensity after the photobleaching correction, in marked contrast
to what we find by treating the constant as a contribution from
unbleachable or regenerated fluorophores.
(a)(b)00.51I(fi; fj)10-2100102(bits)P(I)(1/bits)00.5110-2100102GCaMP6GFPI(fi; fj)..(bits)P(I)(1/bits)4
FIG. 3: Discretization of the empirically observed fluorescence signals. (a) Heatmap of the normalized fluorescence ratio between
photobleaching-corrected GCaMP fluorescence intensity and RFP fluorescence intensity, f , for each neuron as a function of
time. (b) Heatmap of the neuronal activity after discretization based on time derivatives of f . Green corresponds to a state of
"rising", red "falling", and white "flat".
of the trace in Fig. 1d. After the smooth derivative u
is estimated, we discretized the smooth estimate of the
signal, Au, into three states of "rise," "fall," and "flat,"
depending on whether the derivative u exceeds a constant
multiple of σn/τf , the expected standard deviation of the
smooth derivative extracted from a pure white noise. The
constant is chosen to be σn/τf = 5, such that the GFP
control worm has almost all pairwise mutual information
being zero after going through the same data processing
pipeline. An example of the raw fluorescence and final
discretized signals is shown in Fig. 3.
III. MAXIMUM ENTROPY MODEL
After preprocessing, the state of each neuron is de-
scribed by a Potts variable σi, and the state of the entire
network is {σi}. As in previous work on a wide range of
biological systems [11, 14, 17, 36 -- 38], we use a maximum
entropy approach to generate relatively simple approxi-
mations to the distribution of states, P ({σi}), and then
ask how accurate these models are in making predictions
about higher order structure in the network activity.
The maximum entropy approach begins by choosing
some set of observables, Oµ({σi}), over the states of the
system, and we insist that any model we write down for
P ({σi}) match the expectation values for these observ-
ables that we find in the data,
P ({σi})Oµ({σi}) = (cid:104)Oµ({σi})(cid:105)expt.
(3)
(cid:88)
{σi}
Among the infinitely many distributions consistent with
these constraints, we choose the one that has the largest
possible entropy, and hence no structure beyond what is
needed to satisfy the constraints in Eq. (3). The formal
solution to this problem is
P ({σi}) =
1
Z
exp
−
(cid:34)
(cid:35)
λµOµ({σi})
,
(4)
(cid:88)
µ
where coupling constant λµ must be set to satisfy and
Eq. (3), and the partition function Z as usual enforces
normalization.
Following the original application of maximum entropy
methods to neural activity [11], we choose as observables
the mean activity of each cell, and the correlations be-
tween pairs of cells. With neural activity described by
three states, "correlations" could mean a whole matrix
or tensor of joint probabilities for two cells to be in par-
ticular states. We will see that models which match this
tensor have too many parameters to be inferred reliably
from the data sets we have available, and so we take a
simpler view in which "correlation" measures the proba-
bility that two neurons are in the same state. Equation
(4) then becomes
with the effective Hamiltonian
H(σ) = −
1
2
Jijδσiσj −
(cid:88)
i(cid:54)=j
P (σ) =
e−H(σ) ,
(5)
1
Z
(cid:88)
p−1(cid:88)
i
r=1
hr
i δσir .
(6)
The number of states p = 3, corresponding to "rise,"
"fall," and "flat" as defined above. The parameters are
the pairwise interaction Jij and the local fields hr
i , and
these must be set to match the experimental values of
the correlations
cij ≡ (cid:104)δσiσj(cid:105) =
1
T
δσi(t)σj (t) ,
(7)
T(cid:88)
t=1
100200300400500102030405060708010020030040050010203040506070800.511.52100200300400500204060800.511.52Neuron ID,iit(sec)t(sec)(a)(b)risefallflatNormalized fluorescence ratio, fDiscretized signal, σ5
FIG. 5: Top: No signs of overfitting are observed for mod-
els of up to N = 50 neurons, measured by the difference of
per-neuron log-likelihood of the data under the pairwise max-
imum entropy model for training sets consists of 5/6 of the
data and test sets. Clusters around N = 10, 15, 20, . . . , 50
represent randomly chosen subgroups of N neurons. Error
bars are the standard deviation across 10 random partitions
of training and test samples. The dashed lines show the ex-
pected per-neuron log-likelihood difference and its standard
deviation calculated through perturbation methods (see Ap-
pendix A). Bottom: The difference between log likelihood of
the training data and of the test data is greater than 0 (the
red line) within error bars for maximum entropy models on
N = 10, 20, . . . , 50 neurons with pairwise correlation tensor
constraint (see Appendix B), which suggests that this model
does not generalize well.
timate from variations across random halves of the data.
This training procedure leaves part of the interaction ma-
trix Jij zero, while the model is able to reproduce the
magnetization mr
i and the pairwise correlation cij within
the experimental errors (Fig. 4).
Because of the large temporal correlation in the data,
the number of independent data in the recording is small
compared to the number of parameters. This makes us
worry about overfitting, which we test by randomly se-
lecting 5/6 of the data as training set, inferring the max-
imum entropy model from this training set, and then
comparing the log-likelihood of both the training data
and the test data with respect to the maximum entropy
model. No signs of overfitting are found for subgroups of
up to N = 50 neurons, as indicated by that fact that the
difference of the log-likelihood is zero within error bars
(Fig. 5; details in Appendix A). This is not true if we
try to match the full tensor correlations (Appendix B),
which is why we restrict ourselves to the simpler model.
IV. DOES THE MODEL WORK?
The maximum entropy model has many appealing fea-
tures, not least its mathematical equivalence to statisti-
cal physics problems for which we have some intuition.
But this does not mean that this model gives an accu-
FIG. 4: Model construction: learning the maximum entropy
model from data.
(a) Connected pairwise correlation ma-
trix, Cij, measured for a subgroup of 50 neurons. (b) The
inferred interaction matrix, Jij. (c) Probability of neuron i
in state r, for the same group of 50 neurons as panel (a).
(d) The inferred local field, hr
i . (e) Model reproduces pair-
wise correlation (unconnected) within variation throughout
the experiment. Error bars are extrapolated from bootstrap-
ping random halves of the data. (f) Same as panel (e), but
for mean neuron activity mr
i .
and the magnetizations
mr
i ≡ (cid:104)δσir(cid:105) =
1
T
T(cid:88)
t=1
δσi(t)r .
(8)
Note that the local field for the "flat" state, hp
i , is set to
zero by convention. In addition, the interaction Jij can
be non-zero for any pairs of neurons i and j regardless
of the positions of the neurons (both physical and in the
structural connectome), i.e. the equivalent Potts model
does not have a pre-defined spatial structure.
The model parameters are learned using coordinate
descent and Markov chain Monte Carlo (MCMC) sam-
pling [39 -- 41].
In particular, we initialize all parame-
ters at zero. For each optimization step, we calculate
the model prediction cij and mr
i by alternating between
MCMC sampling with 104 MC sweeps and histogram
sampling to speed up the estimation. Then, we choose a
single parameter from the set of parameters {Jij, hr
i} to
update, such that the increase of likelihood of the data is
maximized [39]. We repeat the observable estimation and
parameter update steps until the model reproduces the
constraints within the experimental errors, which we es-
00.51cij, model00.51cij, data00.51mir, model00.51mir, data01020304050Neuron ID00.51mir, data1020304050Neuron ID1020304050Neuron ID-0.500.51020304050Neuron ID1020304050Neuron ID-1011020304050Neuron ID1020304050Neuron ID-0.500.51020304050Neuron ID1020304050Neuron ID-101iijjCijJij01020304050Neuron ID-505hir, modelmirhir(a)(b)(c)(d)iirisefallflatmir, modelmir, datacij, datacij, model(e)(f)N-0.100.1ltest-ltrain1020304050N-0.100.1ltest-ltrainrate description of the real network. Here we test several
predictions of the model. In practice we generate these
predictions by running a long Monte Carlo simulation of
the model, and then treating the samples in this simula-
tion exactly as we do the real data. We emphasize that,
having matched the mean activity and pairwise correla-
tions, there are no free parameters, so that everything
which follows is a prediction and not a fit.
6
p(cid:88)
r=1
Since we use the correlations between pairs of neurons
in constructing our model, the first nontrivial test is to
predict correlations among triplets of neurons,
Cijk =
(cid:104)(δσir − (cid:104)δσir(cid:105))(δσj r − (cid:104)δσj r(cid:105))(δσkr − (cid:104)δσkr(cid:105))(cid:105) .
(9)
More subtly, since we used only the probability of two
neurons being in the same state, we can try to predict
the full matrix of pairwise correlations,
C rs
ij ≡ (cid:104)δσirδσj s(cid:105) − (cid:104)δσir(cid:105)(cid:104)δσj s(cid:105) ;
(10)
note that the trace of this matrix is what we used in
building the model. Scatter plots of observed vs pre-
dicted values for Cijk and C rs
ij are shown in Fig. 6a and
c. In parts b and d of that figure we pool the data, com-
paring the root-mean-square differences between our pre-
dictions and mean observations (model error) with errors
in the measurements themselves. Although not perfect,
model errors are always within 1.5× the measurement
errors, over the full dynamic range of our predictions.
Turning to more global properties of the system, we
consider the probability of k neurons being in the same
state, defined as
(cid:68) p(cid:88)
r=1
(cid:69)
I(cid:80)N
i=1 δσir=k
P (k) ≡
,
(11)
where I is the indicator function.
It is useful to com-
pute this distribution not just from the data, but also
from synthetic data in which we break correlations among
neurons by shifting each cell's sequence of states by an
independent random time. We see in Fig. 7a that the
real distribution is very different from what we would see
with independent neurons, so that in particular the tails
provide a signature of correlations. These data agree very
well with the distributions predicted by the model.
Our model assigns an "energy" to every possible state
of the network [Eq. (6)], which sets the probability of
that state according to the Boltzmann distribution. Be-
cause our samples are limited, we cannot test whether
the energies of individual states are correct, but we can
ask whether the distribution of these assigned energies
across the real states taken on by the network agree with
what it predicted from the model. Figure 7b compares
these distributions, shown cumulatively, and we see that
there is very good overlap between theory and experi-
ment across ∼ 90% of the density, with the data having
a slightly fatter tail than predicted. The good agreement
FIG. 6: Model validation: The model predicts unconstrained
higher order correlations of the data. Panel (a) shows the
comparison between model prediction and data for the con-
nected three-point correlation Cijk for a representative group
of N = 50 neurons. All 19800 possible triplets are plotted
with the blue dot. Error bars are generated by bootstrapping
random halves of the data, and are shown for 20 uniformly
spaced random triplets in red. Panel (b) shows the error
of three-point function ∆Cijk as a function of the connected
three-point function Cijk, binned by its value predicted by the
model, Cijk, model. The red curve is the difference between
data and model prediction. The blue curve is the standard
error from mean of Cijk over the course of the experiment,
extracted by bootstrapping random halves of the experiment.
Panels (c, d) are the same as panels (a, b), but for the con-
nected two-point correlation tensor C rs
ij .
extends over a range of ∆E ∼ 20 in energy, correspond-
ing to predicted probabilities that range over a factor of
exp(∆E) ∼ 108.
The maximum entropy model gives the probability for
the entire network to be in a given state, which means
that we can also compute the conditional probabilities for
the state of one neuron given the state of all the other
neurons in the network. Testing whether we get this right
seems a very direct test of the idea that activity in the
network is collective. This conditional probability can be
written as
P (σi{σj(cid:54)=i}) ∝ exp
gr
i δσir
,
(12)
where the effective fields are combinations of the local
field hr
i and each cell's interaction with the rest of the
network.
(cid:35)
(cid:34)p−1(cid:88)
r=1
gr
i = hr
i +
Jij(δσj r − δσj p) .
(13)
Then the probabilities for the states of neuron i are set
N(cid:88)
j(cid:54)=i
-0.200.2Cijrs, model-0.2-0.100.10.2Cijrs, data-0.100.1Cijrs, model0.0050.010.0150.02Cijrs-0.0100.010.020.03Cijk, model246810Cijk10-3-0.0500.05Cijk, model-0.0500.05ijkC, data(a)(b)(c)(d)datamodelby
V. WHAT DOES THE MODEL TEACH US?
P (σi = r)
P (σi = p)
= egr
i ,
(14)
A. Energy landscape
7
where the last state p is a reference. In Figure 7c and d
we test these predictions. In practice we walk through
the data, and at each moment in time, for each cell, we
compute effective fields. We then find all moments where
the effective field falls into a small bin, and compute the
ratio of probabilities for the states of the one cell, col-
lecting the data as shown. The agreement is excellent,
except at extreme values of the field which are sampled
only very rarely in the data. We note the agreement ex-
tends over a dynamic range of roughly two decades in the
probability ratios.2
FIG. 7: Model validation: comparison between model predic-
tion and data for observables not constrained by the model.
The neuron network has N = 50 neurons. (a) Probability of
k neurons being in the same state. Blue dots are computed
from the data. Yellow dash-dot line is the prediction from a
model where all neurons are independent, generated by ap-
plying a random temporal cyclic permutation to the activity
of each neuron. Purple line is the prediction of the pairwise
maximum entropy model. (b) Tail distribution of the energy
for the data and the model. All error bars in this figure are
extrapolated from bootstrapping. (c, d) Probability ratio of
the state of a single neuron as a function of the effective field
gr
i , binned by the value of the effective field. Error bars are
the standard deviation after binning.
2 The claim that behaviors are collective requires a bit more than
predictability. It is possible that behaviors of individual cells are
predictable from the state of the rest of the network, but that
most of the predictive power comes from interaction with a sin-
gle strongly coupled partner. We have checked that the mutual
information I(σi; gr
i ) is larger than the maximum of I(σi; σk), in
almost all cases.
Maximum entropy models are equivalent to Boltz-
mann distributions and thus define an energy landscape
over the states of the system, as shown schematically in
Fig. 8a.
In our case, as in other neural systems, the
relevant models have interactions with varying signs, al-
lowing the development of frustration and hence a land-
scape with multiple local minima. These local minima
are states of high probability, and serve to divide the
large space of possible states into basins. It is natural to
ask how many of these basins are supported in subnet-
works of different sizes.
To search for energy minima, we performed quenches
from initial conditions corresponding to the states ob-
served in the experiment, as described in [14]. Briefly,
at each update, we change the state of one neuron such
that the decrease of energy is maximized, and we ter-
minate this procedure when no single spin flip will de-
crease the energy; the states that are attracted to local
energy minimum α form a basin of attraction Ωα. As
shown in Fig. 8c, the number of energy minima grows
sub-exponentially as the number of neurons increases.
Note that this approach only gives us the states that the
animal has access to, rather than all metastable states,
whose number is approximated by greedy quench along
a long MCMC trajectory. Nonetheless, the probability
of visiting a basin is similar between the data and the
model, shown by the rank-frequency plot (Fig. 8d).
Whether the energy minima correspond to well defined
collective states depends on the heights of the barriers
between states. Here, we calculate the barrier height
between basins by single-spin-flip MCMC, initialized at
one minimum α and terminating when the state of the
system belongs to a different basin Ωβ; the barrier be-
tween basins Ωα and Ωβ is defined as the maximum en-
ergy along this trajectory. This sampling procedure is
repeated 1000 times for each initial basin to compute the
mean energy barrier. As shown in Fig. 8b, the distribu-
tion of barrier energies strongly overlaps the distribution
of the energy minima, which implies that the minima are
not well separated.
Further visualization of the topography of the energy
landscape is performed by constructing metabasins, fol-
lowing Ref [42]. Here, we construct metabasins by group-
ing the energy minima according to the barrier height;
basins with barrier height lower than a given energy
threshold, ∆E, are grouped into a single metabasin.
This threshold can be varied: at high enough threshold,
the system effectively does not see any local minima; at
low threshold, the partition of the energy landscape ap-
proaches the partition given by the original basins of at-
traction. If the dynamics were just Brownian motion on
the landscape, states within the same metabasin would
transition into one other more rapidly than states be-
longing to different metabasins. As shown in Fig. 8e,
-505girise-4-20246ln(Prise/Pflat)-4-20246gifall-4-20246ln(Pfall/Pflat)-80-60-40-20Energy10-410-210002040k00.050.1P(k)dataind.pairwise(a)(b)(c)(d)modelmodeldatadataprobability8
FIG. 8: Energy landscape of the inferred maximum entropy model. (a) Schematic of the energy landscape with local minima
α, β and the corresponding basin Ωα, Ωβ. Colored in light blue is the metabasin formed at the given energy threshold, ∆E. (b)
Typical distribution of the value of the energy minima and the barriers of a maximum entropy model on N = 30 neurons. The
global energy minimum, E0, is subtracted from the energy, E. (c) The number of energy minima increases sub-exponentially
as number of neurons included in the model increases. Error bars are the standard deviation of 10 different subgroups of N
neurons. (d) The rank-frequency plot for frequency of visiting each basin matches well between data and model for a typical
subgroup of 40 neurons. (e) The number of metabasins, grouped according to the energy barrier, diverges when the energy
threshold ∆E approaches 1 from above.
there is a transition at ∆E ≈ 1.2 from single to multiple
metabasins for all N = 10, 20, and 30. Since the dynam-
ics of the real system do not correspond to a simple walk
on the energy landscape (Appendix C and Fig. 12), we
cannot conclude that this is a true dynamical transition,
but it does suggest that the state space is organized in
ways that are similar to what is seen in systems with such
transitions.
B. Criticality
Maximum entropy models define probability distribu-
tions that are equivalent to equilibrium statistical physics
problems. As these systems become large, we know that
the parameter space separates into distinct phases, sep-
arated by critical surfaces. In several biological systems
that have been analyzed there are signs that these criti-
cal surfaces are not far from the operating points of the
real networks, although the interpretation of this result
remains controversial
[21]. Here we ask simply whether
the same pattern emerges in C. elegans.
One natural slice through the parameter space of mod-
els corresponds to changing the effective temperature of
the system, effectively scaling all terms in the log prob-
ability up and down uniformly. Concretely, we replace
H(σ) → H(σ)/T in Eq (5). We monitor the heat ca-
pacity of the system, as we would in thermodynamics;
here the natural interpretation is of the heat capacity as
being proportional to the variance of the log probability,
so it measures the dynamic range of probabilities that
can be represented by the network. Results are shown in
Fig. 9, for randomly chosen subsets of N = 10, 20, ..., 50
neurons. A peak in heat capacity often signals a critical
point, and here we see that the maximum of the heat
capacity approaches the operational temperature T0 = 1
from below as N becomes larger, suggesting that the full
network is near to criticality.
C. Network topology
The worm C. elegans is special in part because it is
the only organism in which we know (essentially) the full
pattern of connectivity among neurons. Our models also
have a "connectome," since only a small fraction of the
possible pairs of neurons are linked by a nonzero value of
Jij. The current state of our experiments is such that we
αβΩαΩβ10-5100E - E0Emetabasinbasinprobability densitylocal min.barrier10-1101E100102NMBN = 10N=20N=30100rank10-5frequency102100100N100102104Nbasindatamodel20400510152025(a)(c)(d)(e)(b)9
neuron is the all "flat" state, σk = p. Following the
maximum entropy model [Eq. (6)], the probability dis-
tribution for the rest of the network becomes
(cid:101)Pk(σ) ≡ P (σ1, σ2, ...σN−1σk = 3) =
(cid:101)Hk(σ) = −
where the effective Hamiltonian is
Jijδσiσj −
(cid:88)
(cid:88)
1
2
i(cid:54)=j(cid:54)=k
1(cid:101)Zk
e−(cid:101)Hk(σ) , (15)
p−1(cid:88)
(cid:88)
hr
i δσir .
Jikδσip−
i(cid:54)=k
r=1
i(cid:54)=k
(16)
On the other hand, ablation of neuron k means the re-
moval of neuron k from the network, which leads to an
effective Hamiltonian
(cid:98)Hk(σ) = −
1
2
(cid:88)
i(cid:54)=j(cid:54)=k
(cid:88)
p−1(cid:88)
r=1
i(cid:54)=k
Jijδσiσj −
hr
i δσir .
(17)
We examine the effect of clamping and ablation by
Monte Carlo simulation of these modified models. We fo-
cus on the response of individual neurons i to perturbing
neuron k, which is summarized by change in the magne-
tizations, mr
i . But since these also represent the
probabilities of finding the neuron i in each of the states
r = 1, ..., p, we can measure the change as a Kullback --
Leibler divergence,
i → mr
p(cid:88)
r=1
(cid:19)
(cid:18) mr
i
mr
i
DKL =
mr
i log
bits.
(18)
FIG. 10: The topology of the learned maximum entropy
model approaches that of the structural connectome, as the
number of neurons being modeled, N , increases. The two
global topological properties being measured are the cluster-
ing coefficient C (panel (a)) and characteristic path length
L (panel (b)). Here, the inferred network topology for three
different worms is plotted in blue. Red curves are for the
randomized network with the same number of neurons, N ,
and number of connections, NE, as the model, where we ex-
pect Lrandom ∼ ln(N )/ ln(2NE/N ) and Crandom ∼ 2NE/N 2.
The dark blue line corresponds to the network property of
the structural connectome; the dark red line corresponds to
randomized network with number of nodes and edges equal to
those of the structural connectome [43]. Error bars are gen-
erated from the standard deviation across different 10 sub-
groups of N neurons.
FIG. 9: The heat capacity is plotted against temperature for
models with different number of neurons, N . The maximum
of the heat capacity approaches the operational temperature
of the C. elegans neural system T0 = 1 from below as N
increases. Error bars are the standard error across 10 random
subgroups of N neurons.
cannot identify the individual neurons, and so we cannot
check if the effective connectivity in our model is similar
to the anatomical connections. But we can ask statis-
tical questions about the connections, and we focus on
two global properties of the network: the clustering coef-
ficient C, defined as the fraction of actual links compared
to all possible links connecting the neighbors of a given
neuron, averaged over all neurons; and the characteristic
path length L, defined as the average shortest distance
between any pair of neurons. As shown in Fig. 10, the
topology of the inferred networks for all three worms that
we investigated differ from random Erdos-R´enyi graphs
with the same number of nodes (neurons) and links (non-
zero interactions). Moreover, as we increase the num-
ber of neurons that we consider, the clustering coefficient
C and the characteristic path length L approaches that
found in the structural connectome [43].
D. Local perturbation leads to global response
How well can the sparsity of the inferred network ex-
plain the observed globally-distributed pairwise correla-
tion? In particular, we would like to examine the re-
sponse of the network to local perturbations. This test is
of particular interest, since its predictions can be exam-
ined experimentally, as local perturbation of the neural
network can be achieved through optogenetic clamping
or ablation of individual neurons.
The maximum entropy model can be perturbed
through both "clamping" and "ablation." By definition,
the only possible state in which we can clamp a single
10-1100101Temperature01020304050Heat capacityN = 10N = 20N = 30N = 40N = 50model(a)(b)randomstruct.struct. rand.clustering coefficientcharacteristic path lengthNN20406000.20.40.60.812040601.522.53As shown in Fig. 11, the response of the network to the
local perturbation is distributed throughout the network
for both clamping and ablation. However, clamping leads
to much larger DKLs, suggesting that the network is
more sensitive to clamping, and perhaps robust against
(limited) ablation.
Interestingly, this result echoes the
experimental observation that C. elegans locomotion is
easily disturbed through optogenetic manipulation of sin-
gle neurons [44, 45], while ablation of single neurons has
limited effect on the worms' ability to perform different
patterns of locomotion [46 -- 48], although further experi-
mental investigation is needed to test our hypotheses on
network response.
VI. DISCUSSION
Soon it should be possible to record the activity of the
entire nervous system of C. elegans as it engages in rea-
sonably natural behaviors. As these experiments evolve,
we would like to be in a position to ask question about
collective phenomena in this small neural network, per-
haps discovering aspects of these phenomena which are
shared with larger systems, or even (one might hope)
universal. We start modestly, guided by the state of the
data.
We have built maximum entropy models for groups
of up to N = 50 cells, matching the mean activity and
pairwise correlations in these subnetworks. Perhaps our
most important result is that these models work, provid-
ing successful quantitative predictions for many higher
10
order statistical structures in the network activity. This
parallels what has been seen in systems where the neu-
rons generate action potentials, but the C. elegans net-
work operates in a very different regime. The success of
pairwise models in this new context adds urgency to the
question of when and why these models should work, and
when we might expect them to fail.
Beyond the fact that the models make successful quan-
titative predictions, we find other similarities with analy-
ses of vertebrate neural networks. The probability distri-
butions that we infer have multiple peaks, corresponding
to a rough energy landscape, and the parameters of these
models appear close to a critical surface. In addition, we
have shown that the inferred model is sparse, and has
topological properties similar to that of the structural
connectome. Nevertheless, global response is observed
when the modeled network is perturbed locally, in a way
similar to experimental observations.
With the next generation of experiments, we hope to
extend our analysis in four ways. First, longer recording
will allow construction of meaningful models for larger
groups of neurons. If coupled with higher signal -- to -- noise
ratios, it should also be possible to make a more refined
description of the continuous signals relevant to C. el-
egans neurons, rather than having to compress our de-
scription down to a small number of discrete states. Sec-
ond, registration and identification of the observed neu-
rons will make it possible to compare the anatomical con-
nections between neurons with the pattern of interactions
in our probabilistic models. Being able to identify neu-
rons across multiple worms will also allow us to address
the degree of reproducibility across individuals, and per-
haps extend the effective size of data sets by averaging.
Third, optogenetic tools will allow local perturbation of
the neural network experimentally, which can be com-
pared directly with the theoretical predictions in §V.D
above. Finally, improvements in experimental methods
will enable constructions of maximum entropy models for
freely moving worms, with which we can map the relation
between the collective behavior identified in the neuronal
activity and the behavior of the animal.
Acknowledgement
FIG. 11: Local perturbation of the neural network leads to
global response. (a, b) For a typical group of N = 50 neu-
rons, the inferred interaction matrix J is sparse. Here, the
neuron index i and j are sorted based on mflat
, as in Fig. 4.
(c, d) When neuron k is clamped to a constant voltage, the
Kullback-Leibler divergence (in bits) of the marginal distri-
bution of states for neuron i is distributed throughout the
network. (e, f) When neuron k is ablated, the DKL is also
distributed throughout the network, but is smaller than in
response to clamping.
i
We thank F Beroz, AN Linder, L Meshulam, JP
Nguyen, M Scholz, NS Wingreen, and B Xu for many
helpful discussions. Work supported in part by the Na-
tional Science Foundation through the Center for the
Physics of Biological Function (PHY -- 1734030), the Cen-
ter for the Science of Information (CCF -- 0939370), and
PHY -- 1607612; and by the Simons Collaboration on the
Global Brain.
012DKL, ablation10-2100102104-202Jij10-410-2100102probability density-0.500.5S051015012DKL10-2100102(a)(c)(e)-202Jij10-410-2100102probability density-0.500.5S051015012DKL10-2100102DKL, clampingDKL, ablationJijprobability density2040i1020304050j-101Jijji2040i1020304050j-1012040Marginal neuron ID, i1020304050Clamped neuron ID, k00.10.20.30.40.5ik2040Marginal neuron ID, i1020304050Clamped neuron ID, k00.10.20.30.40.5DKLclamping k2040Marginal neuron ID, i1020304050Ablated neuron ID, k00.020.040.060.080.1DKL(PPk)DKLk2040Marginal neuron ID, i1020304050Ablated neuron ID, k00.020.040.060.080.1DKL(PPk)iablatingkprobability densityprobability density(d)(b)(f)11
Now, let us assume that a set of true underlying param-
eters, {g∗}, exists for the system we study, which leads to
a true expectation value be f∗i = fi(g∗). However, we are
only given finite number of observations, σ1, σ2, . . . , σT ,
from which we construct a maximum entropy model, i.e.
infer the parameters {g} by maximizing the likelihood of
the data. Our hope is that the difference between the
true parameters and the inferred parameters is small, in
which case we can approximate the inferred parameters
using a linear approximation
(cid:88)
(21)
where
gi = g∗i + δgi,
∂gi
δgi ≈
∂fj
(cid:101)χijδfj. (22)
Here, (cid:101)χ is the inverse of the susceptibility matrix χij =
δfj = −
−∂fi/∂gj = (cid:104)φiφj(cid:105) − (cid:104)φi(cid:105)(cid:104)φj(cid:105); and δfj is the difference
between empirical mean and the true mean of φj,
j
j
δfj =
1
T
φj(σt) − f∗j
(23)
For convenience, we will use short-hand notation
φi(σt) = φt
i to indicate the value of the observable φi
at time t.
Let the number of samples in the training data be T1,
and the number of samples in the test data be T2. For
simplicity, assume that all samples are independent. We
maximize the entropy of the model on only the training
data to obtain parameters {g}, and we would like to know
how well our model generalize to the test data. Thus,
we quantify the degree of overfitting by the difference of
likelihood of the training data and the test data:
(cid:88)
T(cid:88)
t=1
Author Contributions
XC and WB performed the analyses and the simula-
tions. FR and AML designed and carried out the ex-
periments. All authors contributed to the manuscript
preparation.
Appendix A: Perturbation methods for overfitting
analysis
To test if our maximum entropy model overfits, we
partition the samples into a set of training data and a
set of test data. The difference of the per-neuron log-
likelihood for the training data and the test data is used
as a metric of whether the model overfits: if the two val-
ues for the log-likelihood are equal within error bars, then
the model generalizes well to the test data and does not
overfit. Here, we outline a perturbation analysis which
uses the number of independent samples and the number
of parameters of the model to estimate the expectation
value of this log-likelihood difference.
Consider a Boltzmann distribution parameterized by
g = g1, g2, . . . , gm acting on observables φ1, φ2, . . . , φm.
The probability for the N spins taking the value σ =
σ1, σ2, . . . , σN is
(cid:33)
(cid:32)
m(cid:88)
i=1
P (σg) =
1
Z(g)
exp
−
giφi(σ)
,
(19)
where Z is the partition function. Then, the log-
likelihood of a set of data with T samples under the
Boltzmann distribution parameterized by g is
L(σ1, σ2, . . . , σTg) =
1
T
T(cid:88)
t=1
= − log Z(g) −
gi
i=1
log P (σtg)
m(cid:88)
(cid:32)
(cid:34)
m(cid:88)
i=1
− log Z(g) −
g∗i −
(cid:88)
j
(cid:33)
φt
i
t=1
1
T
T(cid:88)
m(cid:88)
(cid:101)χij
(cid:32)
i=1
gi
(cid:32)
1
T1
Ltest − Ltrain =
=
(cid:34)
(cid:33)(cid:35)
(cid:33)(cid:32)
−
(20)
T2(cid:88)
T1(cid:88)
t(cid:48)=1
1
T2
(cid:48)
φt
i
j − f∗j
φt
t=1
(cid:32)
m(cid:88)
T2(cid:88)
i=1
gi
T1(cid:88)
(cid:33)
t=1
.
1
T1
(cid:48)
φt
i
t(cid:48)=1
− log Z(g) −
T1(cid:88)
t=1
1
T1
φt
i −
1
T2
(cid:33)(cid:35)
φt
i
(24)
For simplicity of notation, let us write
α(1)
i =
1
T1
i − f∗i ,
φt
α(2)
i =
1
T2
T1(cid:88)
t=1
T2(cid:88)
t=1
ference [Eq. (24)], have expectation values
i − f∗i .(25)
φt
(cid:104)α(1)
i
(cid:105) = 0 ,
i α(1)
(cid:104)α((1)
j (cid:105) =
1
T1
χij .
(26)
By the Central Limit Theorem, α(1)
are Gaus-
sian variables. Terms that appear in the likelihood dif-
and α(2)
i
i
In addition, because we assume that the training data
and the test data are independent, the cross-covariance
between the training data and the test data is
Appendix B: Maximum entropy model with the
pairwise correlation tensor constraint
12
(cid:104)Ltest − Ltrain(cid:105) =
i α(2)
(cid:104)α((1)
j (cid:105) = 0 .
(27)
Combining all the above expressions, we obtain the
expectation value of the likelihood difference [Eq. (24)],
(cid:17)(cid:69)
α(1)
i − α(2)
i
(cid:16)
(cid:101)χijα(1)
j
i=1
g∗i −
(cid:68) m(cid:88)
(cid:88)
m(cid:88)
m(cid:88)
(cid:101)χij(cid:104)α(1)
m(cid:88)
m(cid:88)
(cid:101)χijχij
j=1
i=1
j
i α(1)
j (cid:105)
= −
= −
= −
1
T1
m
T1
i=1
j=1
(28)
Note that the difference of likelihood is only related to
the number of parameters in our model and the number
of independent samples in the training data.
Similarly, we can evaluate the variance of the likelihood
difference to be
(cid:104)(Ltest − Ltrain)2(cid:105) =
(cid:18) 1
(cid:19)
g∗i g∗kχik
+
1
T2
T1
(m2 + 2m) +
m
T1T2
(29)
(cid:88)
i,k
+
1
T 2
1
using Wick's theorem for multivariate Gaussian variables
and chain rules of partial derivatives.
In order to test whether perturbation theory can be
applied to the maximum entropy model learned from the
real data, we estimate the number of independent sam-
ples using Nind. sample ∼ T /τ , where T is the length of
the experiment and τ is the correlation time. The cor-
relation time is extracted as the decay exponent of the
overlap function, defined to be
(cid:68) 1
N(cid:88)
N
i=1
(cid:69)
t
q(∆t) =
δσi(t)σi(t+∆t)
,
(30)
In our experiment, the correlation time is τ = 4 ∼ 6s.
For a typical recording of 8 minutes, the number of inde-
pendent samples is between 80 and 120.
In Figure 5, we compute the perturbation results using
the number of non-zero parameters after the training and
the number of independent samples estimated from the
data. The prediction is within the error bar from the
data, which suggests that the inferred coupling is within
the perturbation regime of the true underlying coupling.
Note that the plotted difference is computed for the per-
neuron log-likelihood, ltest − ltrain = (Ltest − Ltrain)/N .
To fully describe the pairwise correlation between neu-
rons with p = 3 states, the equal-state pairwise correla-
tion cij = (cid:104)δσiσj(cid:105) is not enough; rather, we should con-
strain the pairwise correlation tensor, defined as
crs
ij ≡ (cid:104)δσirδσj s(cid:105) .
(31)
s crs
ij = mr
related through normalization requirements,(cid:80)
and (cid:80)
Here, we constrain the pairwise correlation tensor crs
ij to-
gether with the local magnetization mr
i ≡ (cid:104)δσir(cid:105). No-
tice that for each pair of neurons (i, j), the number of
constraints are p2 + 2p = 15, but these constraints are
i = 1
i , which leads to only 7 independent
variables for each pair of neurons. Because of this non-
independence, choosing which variables to constraint is
a problem of gauge fixing. Here, we choose the gauge
where we constrain the local magnetization mr
i for states
ij ≡ crr
"rise" and "fall", and the pairwise correlations cr
ij ;
in this gauge the parameters can be compared meaning-
fully to the equal-state maximum entropy model above.
The corresponding maximum entropy model has the form
r mr
−
1
2
(cid:88)
3(cid:88)
r=1
i(cid:54)=j
(cid:88)
2(cid:88)
i
r=1
P (σ) ∝ exp
J r
ijδσirδσj r −
hr
i δσir
(32)
Note that the equivalence between constraining the
equal-state correlation for each state and constraining the
full pairwise correlation tensor only holds for the case of
p = 3. For p > 3 states, one need to choose more con-
straints to fix the gauge, and it is not obvious which
variables to fix.
We train the maximum entropy model with tensor con-
straint [Eq. (32)] with the same procedure as the model
with equal-state correlation constraint, described in the
main text. The model is able to reproduce the con-
straints with a sparse interaction tensor J. However,
as shown in the bottom panel of Fig. 5, the difference
between ltrain, the per-neuron log likelihood of the train-
ing data (randomly chosen 5/6 of all data) and ltest, the
per-neuron log likelihood of the test data, is greater than
zero within error bars. This indicates that the maxi-
mum entropy model with tensor constraint overfits for
all N = 10, 20, . . . , 50.
Appendix C: Maximum entropy model fails to
predict the dynamics of the neural networks as
expected
By construction, the maximum entropy model is a
static probability model of the observed neuronal activ-
ities. No constraint on the dynamics was imposed in
building the model, and infinitely many dynamical mod-
els can generate the observed static distribution. The
simplest possibility corresponds to the dynamics being
13
like the dynamics of Monte Carlo itself, which is essen-
tially Brownian motion on the energy landscape. To test
whether this equilibrium dynamics can capture the real
neural dynamics of C. elegans, we compare the mean oc-
cupancy time of each basin, (cid:104)τα(cid:105), calculated using the ex-
perimental data and using MCMC. The mean occupancy
time is defined as the average time a trajectory spends in
a basin before escaping to another basin. For equilibrium
dynamics, the mean occupancy time is determined by the
height of energy barriers according to the transition state
theory, or by considering random walks on the energy
landscape, which gives the relation τ ∼ −p2/2e ln(Pα),
where p = 3 is the number of Potts states and Pα is the
fraction of time the system visits basin α. As shown in
Figure 12, the mean occupancy time (cid:104)τ MC
α (cid:105) found in the
Monte Carlo simulation can be predicted by this simple
approximation. In contrast, the empirical neural dynam-
ics deviates from the equilibrium dynamics, as we might
have expected. The dependence between (cid:104)τ data
(cid:105) and
α
0.5±0.027
P data
.
is weak; a linear fit gives (cid:104)τ data
α
α
(cid:105) ≈ P data
α
FIG. 12: Equilibrium dynamics of the inferred pairwise max-
imum entropy model fails to capture the neural dynamics of
C. elegans. The mean occupancy time of each basin of the
energy landscape, (cid:104)τα(cid:105), is plotted against the fraction of time
the system visits the basin, Pα. For 10 subgroups of N = 10
(in dots) and N = 20 (in asterisks) neurons, the empirical
dynamics exhibits a weak power-law relation between (cid:104)τα(cid:105)
and Pα. The striped patterns are artifacts due to finite sam-
ple size. In contrast, equilibrium dynamics extracted from a
Monte Carlo simulation following detailed balance shows an
inverse logarithmic relation between (cid:104)τα(cid:105) and Pα, which can
be explained by random walks on the energy landscape. Error
bars of the data are extracted from random halves of the data.
Error bars of the Monte Carlo simulation are calculated using
correlation time and standard deviation of the observables.
[1] J. J. Hopfield, Proc. Nat. Acad. Sci. USA 79, 2554
Sci. USA 106, 14058 (2009).
(1982).
[13] G. Tkacik, E. Schneidman, I. Berry, J. Michael, and
[2] J. J. Hopfield, Proc. Nat. Acad. Sci. USA 81, 3088
W. Bialek, arXiv preprint arXiv:0912.5409 (2009).
(1984).
[3] D. J. Amit, H. Gutfreund, and H. Sompolinsky, Phys.
Rev. A 32, 1007 (1985).
[14] G. Tkaik, O. Marre, D. Amodei, E. Schneidman,
W. Bialek, and M. J. Berry, II, PLoS Comp. Biol. 10,
1 (2014).
[4] D. A. Dombeck, C. D. Harvey, L. Tian, L. L. Looger, and
[15] R. Monasson and S. Rosay, Phys. Rev. Lett. 115, 098101
D. W. Tank, Nat. Neurosci. 13, 1433 (2010).
(2015).
[5] M. B. Ahrens, M. B. Orger, D. N. Robson, J. M. Li, and
[16] L. Posani, S. Cocco, K. Jezek, and R. Monasson, J.
P. J. Keller, Nat. Methods 10, 413 (2013).
Comp. Neurosci. 43, 17 (2017).
[6] R. Segev, J. Goodhouse, J. Puchalla, and M. J. Berry II,
[17] L. Meshulam, J. L. Gauthier, C. D. Brody, D. W. Tank,
Nat. Neurosci. 7, 1155 (2004).
[7] J. P. Nguyen, F. B. Shipley, A. N. Linder, G. S. Plummer,
M. Liu, S. U. Setru, J. W. Shaevitz, and A. M. Leifer,
Proc. Nat. Acad. Sci. USA 113, E1074 (2016).
[8] J. P. Nguyen, A. N. Linder, G. S. Plummer, J. W. Shae-
vitz, and A. M. Leifer, PLoS Comp. Biol. 13, e1005517
(2017).
[9] V. Venkatachalam, N. Ji, X. Wang, C. Clark, J. K.
Mitchell, M. Klein, C. J. Tabone, J. Florman, H. Ji,
J. Greenwood, et al., Proc. Nat. Acad. Sci. USA 113,
E1082 (2016).
[10] E. T. Jaynes, Phys. Rev. 106, 620 (1957).
[11] E. Schneidman, M. J. Berry II, R. Segev, and W. Bialek,
Nature 440, 1007 EP (2006).
[12] S. Cocco, S. Leibler, and R. Monasson, Proc. Natl. Acad.
and W. Bialek, Neuron 96, 1178 (2017).
[18] A. Tang, D. Jackson, J. Hobbs, W. Chen, J. L. Smith,
H. Patel, A. Prieto, D. Petrusca, M. I. Grivich, A. Sher,
et al., J. Neurosci. 28, 505 (2008).
[19] I. E. Ohiorhenuan, F. Mechler, K. P. Purpura, A. M.
Schmid, Q. Hu, and J. D. Victor, Nature 466, 617 (2010).
[20] U. Koster, J. Sohl-Dickstein, C. M. Gray, and B. A. Ol-
shausen, PLoS Comp. Biol. 10, e1003684 (2014).
[21] T. Mora and W. Bialek, J. Stat. Phys. 144, 268 (2011).
[22] M. A. Munoz, Rev. Mod. Phys. 90, 031001 (2018).
[23] L. Meshulam, J. L. Gauthier, C. D. Brody, D. W.
Tank, and W. Bialek, arXiv preprint arXiv:1809.08461
[q-bio.NC] (2018).
[24] F. Rieke, D. Warland, R. de Ruyter van Steveninck,
and W. Bialek, Spikes: Exploring the Neural Code (MIT
10-410-310-210-1100P10-1100101102 (sec)datamodelh⌧data↵i/P0.5±0.027↵<latexit sha1_base64="i6SKBv5WghJZpVb2NiJZSL/H21w=">AAACLXicbVDLSiNBFK12fMZXHJduCoPgKnSLEpeCLmYZwUQhHcPtyk1SWF1dVN0WQ5MfcuOvDAOziIjb+Y2pPARfBwoO59zLrXMSo6SjMBwHCz8Wl5ZXVtdK6xubW9vlnZ9Nl+VWYENkKrM3CThUUmODJCm8MRYhTRReJ3fnE//6Hq2Tmb6iocF2Cn0te1IAealTvogV6L7CmCDvxKDMAG5jwgcqukAwiu3MNTYzlPH620gRVk9ik/KwGh7VRp1yxZMp+FcSzUmFzVHvlP/E3UzkKWoSCpxrRaGhdgGWpFA4KsW5QwPiDvrY8lRDiq5dTNOO+IFXuryXWf808an6fqOA1LlhmvjJFGjgPnsT8TuvlVPvtF1IbXJCLWaHerniPvikOt6VFgWpoScgrPR/5WIAFgT5gku+hOhz5K+keVSNwmp0eVw5C+d1rLI9ts8OWcRq7Iz9YnXWYII9st9szJ6Dp+Bv8BK8zkYXgvnOLvuA4N9/oLqoxw==</latexit><latexit sha1_base64="i6SKBv5WghJZpVb2NiJZSL/H21w=">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</latexit><latexit sha1_base64="i6SKBv5WghJZpVb2NiJZSL/H21w=">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</latexit><latexit sha1_base64="i6SKBv5WghJZpVb2NiJZSL/H21w=">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</latexit>h⌧RW↵i= 92e1logP↵<latexit sha1_base64="4SJrWOe/r0wKavyxMei7GxFDvlc=">AAACM3icbVDPSxtBGJ211mpsNa1HL4NB6KVhVwq1B0HwIj2l0hghk4ZvJ98mg7Ozy8y30rDs/+TFf8SDUDy0FK/+D06SPdQfDwYe732Pb74X51o5CsPfwdKr5dcrb1bXGutv321sNt9/OHVZYSV2ZaYzexaDQ60MdkmRxrPcIqSxxl58fjTzexdoncrMD5rmOEhhbFSiJJCXhs1vQoMZaxQExVCAzifwUxD+ovKkVwk79/gB/yQSC7L8WpV7WC14VJVCZ2PeqWPVsNkK2+Ec/DmJatJiNTrD5rUYZbJI0ZDU4Fw/CnMalGBJSY1VQxQOc5DnMMa+pwZSdINyfnPFd70y4klm/TPE5+r/iRJS56Zp7CdToIl76s3El7x+Qcn+oFQmLwiNXCxKCs0p47MC+UhZlKSnnoC0yv+Vywn4RsjX3PAlRE9Pfk5O99pR2I6+f24dhnUdq2yb7bCPLGJf2CE7Zh3WZZJdshv2h/0NroLb4F9wtxhdCurMFnuE4P4BgC+r8g==</latexit><latexit sha1_base64="4SJrWOe/r0wKavyxMei7GxFDvlc=">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</latexit><latexit sha1_base64="4SJrWOe/r0wKavyxMei7GxFDvlc=">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</latexit><latexit sha1_base64="4SJrWOe/r0wKavyxMei7GxFDvlc=">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</latexit>(sec)h⌧data↵i<latexit sha1_base64="XOb/o3matVuyZZFJ2VbX83Rdrqw=">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</latexit><latexit sha1_base64="XOb/o3matVuyZZFJ2VbX83Rdrqw=">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</latexit><latexit sha1_base64="XOb/o3matVuyZZFJ2VbX83Rdrqw=">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</latexit><latexit sha1_base64="XOb/o3matVuyZZFJ2VbX83Rdrqw=">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</latexit>(MC sweep)h⌧MC↵i<latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="hP+6LrUf2d3tZaldqaQQvEKMXyw=">AAAB2XicbZDNSgMxFIXv1L86Vq1rN8EiuCozbnQpuHFZwbZCO5RM5k4bmskMyR2hDH0BF25EfC93vo3pz0JbDwQ+zknIvSculLQUBN9ebWd3b/+gfugfNfzjk9Nmo2fz0gjsilzl5jnmFpXU2CVJCp8LgzyLFfbj6f0i77+gsTLXTzQrMMr4WMtUCk7O6oyaraAdLMW2IVxDC9YaNb+GSS7KDDUJxa0dhEFBUcUNSaFw7g9LiwUXUz7GgUPNM7RRtRxzzi6dk7A0N+5oYkv394uKZ9bOstjdzDhN7Ga2MP/LBiWlt1EldVESarH6KC0Vo5wtdmaJNChIzRxwYaSblYkJN1yQa8Z3HYSbG29D77odBu3wMYA6nMMFXEEIN3AHD9CBLghI4BXevYn35n2suqp569LO4I+8zx84xIo4</latexit><latexit sha1_base64="2O57DSZK9mfx3RPtw5h2brH+nsQ=">AAACA3icbZC9SgNBFIXv+m/8i7Y2gyJYhV0bLQUbG0HBmEA2hruTm2RwdnaZuSuGJU9g46vYWCjiO9j5Nk5iCv8ODBzOmeHO/ZJcK8dh+BHMzM7NLywuLVdWVtfWN6qbq1cuK6ykusx0ZpsJOtLKUJ0Va2rmljBNNDWSm5Nx37gl61RmLnmYUzvFvlE9JZF91KnuxRpNX5OIGYtOjDof4HXMdMfl2clIxHbSdqq7YS2cSPw10dTswlTnnep73M1kkZJhqdG5VhTm3C7RspKaRpW4cJSjvME+tbw1mJJrl5N1RmLPJ13Ry6w/hsUk/f6ixNS5YZr4mynywP3uxuF/Xavg3lG7VCYvmIz8GtQrtOBMjNmIrrIkWQ+9QWmV/6uQA7Qo2ROseAjR75X/mquDWhTWoosQlmAbdmAfIjiEYziFc6iDhHt4hGd4CR6Cp+D1C9dMMOW2BT8UvH0C3uea9g==</latexit><latexit sha1_base64="PQSt19o7XeD8eef3KCoKuz3z6nI=">AAACA3icdZDNSiNBFIVv64w60dGM29kUI4Krpjv+JO4G3MxGcMCokI7hduUmKayubqpui6HJE7jxVWbjQhnmHWY3b2MlKjiiBwoO51Rx635poZXjKPoXzM1/+LiwuPSptrzyeXWt/mXlxOWlldSWuc7tWYqOtDLUZsWazgpLmKWaTtOLg2l/eknWqdwc87igboZDowZKIvuoV99MNJqhJpEwlr0EdTHC84TpiqvDg4lI7Kzt1TeicL+xu7e7I6Jwpxlvt7a9acTNViMWcRjNtAFPOurV/yb9XJYZGZYanevEUcHdCi0rqWlSS0pHBcoLHFLHW4MZuW41W2ciNn3SF4Pc+mNYzNKXLyrMnBtnqb+ZIY/c624avtV1Sh60upUyRclk5OOgQakF52LKRvSVJcl67A1Kq/xfhRyhRcmeYM1DeN5UvG9OGmEchfHPCJbgK3yDLYihCd/hBxxBGyRcwy+4g/vgJrgNfj/imgueuK3Dfwr+PABM4ZtE</latexit><latexit sha1_base64="w5CfhoMF57sszR/7ktr2I43mCzs=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">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</latexit><latexit sha1_base64="ytC6xhkR+dl6rAIVagyF1ctAOPw=">AAACDnicdVDBSiNBEO1Rd3WjuxvXo5fGIHgaZmLcxJuQixchglEhE0NNp5I09vQM3TViGPIFXvZXvHhQlr169rZ/YydGcBd9UPB4r4qqenGmpKUg+OstLC59+ry88qW0uvb12/fy+o9Tm+ZGYFukKjXnMVhUUmObJCk8zwxCEis8iy+bU//sCo2VqT6hcYbdBIZaDqQAclKvvB0p0EOFPCLIexGobAQXEeE1FUfNCY/MzO2VK4G/X937uVfjgV+rh7uNXUeqYb1RDXnoBzNU2BytXvkp6qciT1CTUGBtJwwy6hZgSAqFk1KUW8xAXMIQO45qSNB2i9k7E77tlD4fpMaVJj5T304UkFg7TmLXmQCN7P/eVHzP6+Q0aHQLqbOcUIuXRYNccUr5NBvelwYFqbEjIIx0t3IxAgOCXIIlF8Lrp/xjclr1w8APj2uVg2AexwrbZFtsh4Wszg7YIWuxNhPsht2ye/bg/fLuvN/en5fWBW8+s8H+gff4DAKjnKs=</latexit>Probability of visiting basin α, PαMean occupancy time14
Press, 1997).
[25] M. B. Goodman, D. H. Hall, L. Avery, and S. R. Lockery,
Neuron 20, 763 (1998).
[26] G. J. Stephens, M. Bueno de Mesquita, W. S. Ryu, and
W. Bialek, Proc. Nat. Acad. Sci. USA 108, 7286 (2011).
[27] P. Sengupta and A. D. Samuel, Curr. Opin. Neurobiol.
19, 637 (2009).
[28] E. L. Ardiel and C. H. Rankin, Learn. Mem. 17, 191
(2010).
Proc. Nat. Acad. Sci. USA 107, 5405 (2010).
[38] W. Bialek, A. Cavagna, I. Giardina, T. Mora, E. Silvestri,
M. Viale, and A. M. Walczak, Proc. Nat. Acad. Sci. USA
109, 4786 (2012).
[39] M. Dud´ık, S. J. Phillips, and R. E. Schapire, in Pro-
ceedings of the 17th annual Conference on Learning The-
ory,(COLT 2004), Banff, Canada (Springer, 2004), vol.
3120, pp. 472 -- 486.
[40] T. Broderick, M. Dudik, G. Tkacik, R. E. Schapire, and
[29] A. L. Nichols, T. Eichler, R. Latham, and M. Zimmer,
W. Bialek, arXiv preprint arXiv:0712.2437 (2007).
Science 356, eaam6851 (2017).
[30] T. Bullock and G. A. Horridge, Structure and function
in the nervous systems of invertebrates. (San Francisco,
1965).
[31] J. G. White, E. Southgate, J. N. Thomson, and S. Bren-
ner, Philos. Trans. R. Soc. Lond. B Biol. Sci. 314, 1
(1986).
[32] T.-W. Chen, T. J. Wardill, Y. Sun, S. R. Pulver, S. L.
Renninger, A. Baohan, E. R. Schreiter, R. A. Kerr, M. B.
Orger, V. Jayaraman, et al., Nature 499, 295 (2013).
[41] M. Schmidt, UGM: A Matlab toolbox for probabilis-
tic undirected graphical models, http://www.cs.ubc.ca/
~schmidtm/Software/UGM.html (2007).
[42] O. M. Becker and M. Karplus, J. Chem. Phys. 106, 1495
(1997).
[43] D. J. Watts and S. H. Strogatz, Nature 393, 440 (1998).
[44] A. Gordus, N. Pokala, S. Levy, S. W. Flavell, and C. I.
Bargmann, Cell 161, 215 (2015).
[45] M. Liu, A. K. Sharma, J. Shaevitz, and A. M. Leifer,
eLife 7, e36419 (2018).
[33] E. Kim, L. Sun, C. V. Gabel, and C. Fang-Yen, PLoS
[46] J. M. Gray, J. J. Hill, and C. I. Bargmann, Proc. Nat.
One 8, e53419 (2013).
Acad. Sci. USA 102, 3184 (2005).
[34] N. Slonim, G. S. Atwal, G. Tkacik, and W. Bialek, arXiv
[47] B. J. Piggott, J. Liu, Z. Feng, S. A. Wescott, and X. S.
preprint cs/0502017 (2005).
Xu, Cell 147, 922 (2011).
[35] R. Chartrand, ISRN Appl. Math. 2011 (2011).
[36] M. Weigt, R. A. White, H. Szurmant, J. A. Hoch, and
T. Hwa, Proc. Nat. Acad. Sci. USA 106, 67 (2009).
[37] T. Mora, A. M. Walczak, W. Bialek, and C. G. Callan,
[48] G. Yan, P. E. V´ertes, E. K. Towlson, Y. L. Chew, D. S.
Walker, W. R. Schafer, and A.-L. Barab´asi, Nature 550,
519 (2017).
|
1103.2974 | 1 | 1103 | 2011-03-15T17:49:45 | Near-Infrared Fluorescence Enhanced (NIR-FE) Molecular Imaging of Live Cells on Gold Substrates | [
"physics.bio-ph"
] | Low quantum yields of near infrared (NIR) fluorophores have limited their capabilities as imaging probes in a transparent, low background imaging window. Here for the first time we reported near-infrared fluorescence enhance (NIR-FE) cell imaging using nanostructured Au substrate, which was employed as a general platform for both single-walled carbon nanotubes (SWNTs) and organic fluorescent labels in the NIR region. Fluorescence intensity, as well as cell targeting specificity, was greatly improved by this novel imaging technique. With NIR-FE imaging, we were able to image SWNT-stained cells at short exposure time of 300ms, and push the detectable limit of SWNT staining of cells down to an ultralow concentration of ~50 pM. Further, different degrees of fluorescence enhancement for endocytosed, intracellular SWNTs vs. nanotubes on the cell membrane at the cell/gold interface were observed, suggesting the possibility of using this technique to track the transmembrane behavior of NIR fluorophores. | physics.bio-ph | physics | 1
Near-Infrared Fluorescence Enhanced (NIR-FE) Molecular Imaging of Live
Cells on Gold Substrates**
Guosong Hong, Scott M. Tabakman, Kevin Welsher, Zhuo Chen, Joshua T. Robinson,
Hailiang Wang, Bo Zhang and Hongjie Dai
Prof. H. Dai, G. Hong, S. M. Tabakman, J. T. Robinson, H. Wang, B. Zhang
Department of Chemistry, Stanford University, Stanford, California 94305
E-mail: [email protected]
Dr. K. Welsher
Department of Chemistry, Princeton University, Princeton, New Jersey, 08544
Dr. Z. Chen
State Key Laboratory of Chemo/Biosensing and Chemometrics, College of Biology,
Hunan University, Changsha, Hunan, 410082 (China)
Biological imaging in the near infrared (NIR) window enjoys advantages of deep
photon penetration due to relatively high transparency, reduced scattering and low
autofluorescence of biological tissues in the 0.8-1.4 μm range.[1-5] Organic dyes[6-9],
quantum dots[10-12] and single-walled carbon nanotubes (SWNTs)[4] have been
employed for in vitro and in vivo biological imaging in the NIR region. SWNTs are a
group of one-dimensional (1-D) macromolecular fluorophores, with intrinsic bandgap
fluorescence emission between 0.9-1.4 μm upon excitation in the visible or NIR.[4,
13-14] The large Stokes shift makes SWNTs ideal probes for biological imaging with
high contrast and low background.[3] Thus far, SWNTs have been used as in vitro
fluorescence tags for cell imaging,[15-16] ex vivo imaging of tissues and organs,[17-18]
and in vivo imaging of normal organs as well as tumors.[19-20]
A common caveat of NIR fluorophores is the relatively low quantum yields
compared to their counterparts (including organic dyes and quantum dots) with
shorter emission wavelengths in the visible, which limits their imaging capabilities.
For example, the IR800 dye (with a peak emission wavelength of 800 nm) exhibits a
~10% quantum yield,[21] and the indocyanine green (ICG) dye exhibits only ~4.3%
quantum yield at the emission wavelength of 805 nm.[22] In contrast, molecules
fluorescing at shorter wavelengths typically exhibit much higher quantum yields
(IR700 ~24% at 700 nm emission[21]; cyanine-5 ~30% at 660 nm emission[23];
fluorescein ~91% at 521 nm emission[24]). SWNTs exhibit quantum yield ranging
2
from 0.1% to 3%,[13, 25-26] due to intrinsic low-energy excitons that are optically
forbidden,[27] and extrinsic quenchers such as metallic SWNTs in bundles[25] and
oxygen in acidic environment[28-29]. To fully utilize the spectral advantages of NIR
fluorophores, it is desirable to develop a general approach to enhancing the
photoluminescence (PL) in the NIR, thus enhancing the biological imaging capability
in this important spectral region.
Recently, we reported metal-enhanced fluorescence (MEF) by up to 14-fold for
surfactant-coated SWNTs placed on top of nanostructured gold films synthesized
purely in the solution phase (called ‘Au/Au films’).[30] Here, we demonstrated this
Au/Au substrate as a powerful and general platform for NIR fluorescence enhanced
(NIR-FE) cellular imaging using both SWNT and organic fluorescent labels. We used
SWNTs functionalized by arginine–glycine–aspartic acid (RGD) to selectively tag
U87-MG brain cancer cells over MCF-7 breast cancer cells, plated the cells on the
Au/Au substrate, and observed a ~9-fold increase in SWNT fluorescence on U87-MG
cells. This enabled high quality NIR molecular imaging of molecularly targeted cells
using much shorter exposure times (~300 ms) than previously possible with nanotube
fluorophores. With NIR-FE imaging, we were able to push the detectable limit of
SWNT staining of cells down to an ultralow concentration of ~50 pM. Further, we
observed different degrees of fluorescence enhancement for endocytosed, intracellular
SWNTs vs. nanotubes on the cell membrane at the cell/gold interface, suggesting the
possibility of observing transmembrane endocytosis of live cells based on the degree
of fluorescence enhancement.
Also important is that our NIR-FE imaging of biological system is general for
commonly used low quantum yield organic dyes including IR800. To our knowledge,
this is the first fluorescence enhanced imaging of cells on Au nanostructures in the
NIR. Previously, Ag substrates were used for fluorescence enhanced biological
imaging in the visible with organic dyes.[31-33]
Cell-type selective staining and subsequent imaging of cells were carried out with
RGD and IR800 conjugated SWNTs, water-solubilized by 25% DSPE-PEG(5k)-NH2
(1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino(polyethyleneglycol)
3
5,000]) and 75% C18-PMH-mPEG(90k) (poly(maleic anhydride-alt-1-octadecene)
-methoxy(polyethyleneglycol)90,000). The RGD peptide ligand was linked to the
amine groups on SWNTs for selectively binding to αvβ3-integrin positive U87-MG
cells[34] over the αvβ3-integrin negative MCF-7 cells. Meanwhile, we also covalently
attached IR800 dye molecules onto SWNTs to afford SWNT-IR800-RGD conjugates
(Figure 1a). Upon excitation at 658 nm, the SWNT-IR800-RGD conjugate emitted in
the range of 1000-1400 nm, due to the intrinsic bandgap photoluminescence of
SWNTs (Figure 1e and 1f). Upon excitation at 785 nm, the conjugate emitted
fluorescence in the 800-1000 nm range due to the attached IR800 molecules (emission
tail also seen in Figure 1e and 1f). This unique SWNT-IR800 conjugate allowed for
tagging of cells using two NIR fluorophores in two different imaging windows in the
800-1400nm range. Atomic force microscopy (AFM) imaging (Figure 1b) showed the
SWNT conjugates with lengths ranging from 100 nm to 3 μm and an average length
of ~1.0 μm.
We synthesized Au/Au films on quartz (Figure 1c) via solution phase growth, with
optimal optical extinction in the NIR (Figure 1d) for the highest fluorescence
enhancement of both SWNTs and IR800 placed on top of the gold film.[30, 35] For
SWNT-IR800-RGD conjugates drop-dried from a solution onto both bare quartz and
Au/Au film on quartz, photoluminescence versus excitation (PLE) spectra revealed
fluorescence enhancement of both IR800 (only emission tail showing up at the upper
left corner of Figure 1e and 1f in the spectral range) and SWNTs (all other peaks in
the 1.0-1.4 μm emission window of Figure 1e and 1f) on the Au/Au film. The average
enhancement of SWNT photoluminescence was ~10 times, and the enhancement was
~5 times for the IR800 dye attached to SWNTs. This result clearly showed the
excellent capability of fluorescence enhancement by the Au/Au film for fluorophores
emitting in the 0.8-1.4 μm NIR window including IR800 and SWNTs.
4
a)
RGD
RGD
Emission:
1000-1400 nm
1000-1400nm
Emission:
800-1000 nm
b)
(a)
IR800
IR800
c)
]
m
n
[
n
o
i
t
a
t
i
c
x
E
e)
5mm d)
D
O
/
n
o
i
t
c
n
i
t
x
E
4.0
3.5
3.0
2.5
2.0
1.5
1.0
400
IR800
800
1200
Wavelength/nm
1600
f)
Emission [nm]
Figure 1. a) A schematic drawing of the SWNT-IR800-RGD conjugate, with the
emission ranges of SWNTs and IR800 dye labeled. b) An AFM image of
SWNT-IR800-RGD conjugates deposited on a SiO2 substrate. c) A digital photograph
of a typical Au/Au substrate used for NIR-FE imaging of cells plated on this substrate.
Inset showed an SEM image of the film. d) A UV-Vis-NIR extinction spectrum of the
Au/Au film. e) A PLE spectrum of SWNT-IR800-RGD conjugates deposited on
quartz. f) A PLE spectrum of SWNT-IR800-RGD conjugates deposited on top of an
Au/Au film.
For targeted cell staining and imaging, U87-MG cells and MCF-7 cells were
trypsinized and mixed with SWNT-IR800-RGD conjugates at a concentration of ~30
nM of SWNTs at 4 ºC for 1h to prevent endocytosis during staining. The cells were
split into two groups and placed onto a quartz microscope slide and Au/Au film
respectively for immediate fluorescence imaging using an InGaAs camera. The
5
αvβ3-integrin positive U87-MG cells treated with the SWNT-IR800-RGD conjugate
showed ~9-fold higher SWNT fluorescence signal (green) on Au/Au (Figure 2a) than
on quartz (Figure 2b), excited at 658 nm under a short exposure time of ~300 ms.
Much
to obtain high quality
times (1~3s) were needed
longer exposure
SWNT-stained cell images on quartz, similar to previous biological imaging with
SWNT fluorophores[19, 36]. The αvβ3-integrin negative MCF-7 cells on both Au/Au
(Figure 2c) and quartz (Figure 2d) showed little SWNT fluorescence signal. The
selectivity of RGD-SWNT labeling of cells, defined as the ratio of SWNT emission
intensity of αvβ3-integrin positive U87-MG cells compared to that of αvβ3-integrin
negative MCF-7 cells, was as high as ~17 for cells on Au/Au substrate (Figure 2g,
higher than the positive/negative ratio of ~7 on quartz), suggesting highly selective
staining and molecular imaging of cells with NIR-FE on the gold films. Note that the
cells appeared round-shaped since they were imaged immediately after being placed
on the substrates prior to adhesion. To show that these imaged cells were alive during
and after imaging, we monitored the U87-MG cells in cell medium after increasing
the temperature to 37 ºC in a temperature controlled imaging chamber with a 1 L/min
CO2 gas flow over a period of 6 hours after the first imaging. Both fluorescence image
and optical image showed cells adhesion to the Au/Au surface at 37 ºC, suggesting
live cells.
We also trypsinized and mixed cells with SWNT-IR800-RGD conjugates at 37 ºC
(instead of 4 ºC as above) for 1h, a condition known to afford endocytosis of carbon
nanotubes inside cells.[37-39] Single particle tracking of SWNTs in live cells has been
studied by Strano and coworkers to reveal the mechanism of endocytosis at 37
ºC.[40-41] The U87-MG and MCF-7 cells thus treated were plated onto Au/Au and
quartz for NIR imaging. In contrast to the ~9-fold enhancement observed for cells
stained at 4 ºC, nanotube fluorescence in the αvβ3-integrin positive U87-MG cells
treated at 37 ºC was enhanced by only ~2-fold on Au/Au film (Figure 2h) compared
to on quartz substrate (Figure 2i). Also noticeable was the higher false-positive signal
intensity in the αvβ3-integrin negative MCF-7 cells (Figure 2j, 2k), due to the expected
increase of non-specific uptake of SWNTs by cells at 37 ºC than at 4 ºC.
on Au/Au
a)
30µm
on quartz
b)
4x
c)
d)
6
4ºC
f)
Au/Au film
quartz
e)
g)
on Au/Au
h)
30µm
i)
on quartz
l)
37ºC
m)
j)
k)
Au/Au film
quartz
n)
)
+
(
G
M
-
7
8
U
)
-
(
7
-
F
C
M
)
+
(
G
M
-
7
8
U
)
-
(
7
-
F
C
M
Figure 2. NIR fluorescence images of a) SWNT-IR800-RGD stained U87-MG
cells at 4ºC on Au/Au film, b) on quartz, and similarly treated MCF-7 cells c) on
Au/Au and d) on quartz. The schematic drawings showed e) SWNTs sandwiched
between cell membrane and gold were enhanced, compared to f) SWNT stained cell
on quartz. g) A bar chart showing average cell fluorescence in each case of 2a)-d).
NIR fluorescence images of h) SWNT-IR800-RGD stained U87-MG cells at 37ºC on
Au/Au film, i) on quartz, and similarly treated MCF-7 cells j) on Au/Au and k) on
quartz. The schematic drawings showed l) SWNTs endocytosed by the cell were
hardly enhanced by gold, compared to m) SWNT stained cell on quartz due to
increased distance between the fluorophores and the enhancing gold substrate. n) A
bar chart showing average cell fluorescence in each case of 2h)-k). All false-colored
7
fluorescence images were taken under a 658 nm excitation with photons collected in
the 1100-1700 nm region.
We attributed the fluorescence enhancement of the molecularly selective SWNT
labels on U87-MG cells to coupling between the emissions of SWNT tags and surface
plasmon modes in the Au/Au substrates. Resonance coupling between SWNT
emission and re-radiating plasmonic modes in the Au/Au films shortened the radiative
lifetimes of SWNTs, affording higher fluorescence quantum efficiency.[30, 42-43] It was
found that surfactant-coated SWNTs closer to the Au/Au surface exhibited higher
fluorescence enhancement, decaying when SWNTs were placed away from the
surface with a half-decay distance of ~5 nm,[30] on the same order of cell membrane
thickness.[44] At 4 ºC, most of the SWNTs were blocked from endocytotic uptake by
the U87-MG cells, and SWNTs on the cell membrane interfacing with the Au/Au
substrate were strongly coupled to the surface plasmonic modes in the gold film and
thus responsible for the large, ~9-fold enhancement in fluorescence compared to on
quartz (Figure 2e, 2f). On the other hand, when incubated at 37 ºC, SWNTs were
endocytosed into the cytoplasms of the cells and hence spatially separated from the
gold surface, giving a reduced fluorescence enhancement of ~2-fold (Figure 2l, 2m).
Also interesting was that we monitored the U87-MG cells stained by SWNTs at 4°C
in cell medium after increasing the temperature to 37 ºC in a 1 L/min CO2 gas flow.
We observed a ~ 6-fold decrease of SWNT fluorescence intensity in the U87-MG
cells over time, from the fluorescence intensity taken right after 4 ºC staining. This
decrease was due to transmembrane endocytosis of SWNTs at 37 ºC, which reduced
the NIR-FE effect as nanotubes were further away from the Au/Au surface. This also
confirmed that the cells were fully alive and functioning.
The SWNT-Au distance dependent fluorescence enhancement could also explain
the measured increase in cellular targeting selectivity with cells on Au/Au vs. quartz
substrate (Figure 2g). For integrin negative MCF-7 cells, the fluorescence signals
detected were due to autofluorescence inside the cells and non-specific uptake effects,
which were distributed through the cells in three dimensions. These non-specific
8
signals were barely enhanced by the Au/Au substrate, while the specific SWNT
signals on the integrin-positive U87-MG cells at the cell-gold interface were enhanced
to the maximum degree due to proximity to Au. Thus, preferential enhancement of
specific cell membrane surface fluorescence afforded more sensitive and selective
imaging of cell membrane receptors. This effect was consistent with little
enhancement in the cell labeling selectivity observed on Au/Au substrate for cells
treated by SWNT-RGD at 37 ºC (Figure 2n). Interestingly, these results suggested
that the distance dependent fluorescence enhancement effect could be used for
tracking transmembrane behavior in live cells, since the thickness of cell membrane
was on the same order of magnitude as the enhancement decay distance (~5 nm).
Besides the distance dependence of fluorescence enhancement, another reason for
the observed increase in targeting selectivity on Au/Au film could be the nonlinearity
of the enhancement effect. It was known that surface enhanced Raman scattering was
non-linear to concentration, i.e., higher concentrations of analytes were usually
enhanced more due to a better chance to occupy the enhancing ‘hot spots’.[45-46]
Therefore, the negative MCF-7 cells were not enhanced as much due to much fewer
SWNTs on the membranes than the positive U87-MG cells, and as a result, targeting
selectivity was magnified by this nonlinear effect.
Our NIR-FE imaging of cells was general for various NIR fluorescent organic dyes.
We chose IR800 as a representative organic dye due to its wide use for biological
imaging.[47-48] The IR800 dye molecules (shown as red circle in Figure 1a) bound to
SWNTs deposited on the same Au/Au film exhibited a fluorescence enhancement by
~5-fold (Figure 1e vs 1f). We used SWNT-IR800-RGD conjugates to target
αvβ3-integrin positive U87-MG cells and performed cell fluorescence imaging in the
IR800 fluorescence channel. Comparing Figure 3a and 3c corresponding to
SWNT-IR800-RGD stained U87-MG and MCF-7 cell lines respectively, one sees that
the αvβ3-integrin positive U87-MG cells showed positive IR800 fluorescence signal
(red) in the 790-820 nm region upon 785 nm excitation, while the αvβ3-integrin
negative MCF-7 cells showed little IR800 signal under the same imaging condition.
This again confirmed high specificity of molecular imaging and the coexistence of
9
IR800 and RGD on SWNTs. Comparison of Figure 3a and 3b revealed a significant
fluorescence enhancement on Au/Au vs. quartz by ~6-fold for IR800 labels on cells,
with a positive/negative selectivity ratio of ~16 on Au/Au film vs. ~4 on quartz
(Figure 3e). These results demonstrated the generality of NIR-FE imaging of cells for
high molecular sensitivity and selectivity.
on quartz
on Au/Au
a)
b)
)
+
(
G
M
-
7
8
U
)
-
(
7
-
F
C
M
10µm
c)
d)
e)
l
l
e
e
c
C
n
d
e
c
e
s
z
e
i
l
a
r
o
m
u
r
l
o
F
N
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
U87-MG cell line
MCF-7 cell line
on AuAu on quartz
Figure 3. IR800 fluorescence images of SWNT-IR800-RGD stained U87-MG cells
on a) Au/Au film and b) on quartz, and similarly treated MCF-7 cells on c) Au/Au
and d) on quartz. Incubation was 1 h for both cell lines at 4 ºC. False-colored images
were taken under 785 nm excitation with photons collected in the 790-820 nm region.
The bar chart in e) showed normalized cell fluorescence in each case of 3a)-d).
Compared to organic NIR fluorophores, surfactant-solubilized SWNTs exhibit even
lower quantum yield ranging from 0.1% to 3%.[13, 25-26] Relatively high concentrations
of SWNTs (~60 nM) are typically needed for biological imaging, where the integrin
binding sites to RGD on cell membranes were almost saturated.[16] To investigate the
detection limit of NIR-FE cell imaging on gold, we labeled αvβ3-integrin positive
U87-MG cells using different staining concentrations of SWNTs from 30 nM to 48
pM and imaged the cells on both Au/Au film (Figure 4, top row) and quartz slide
10
(Figure 4, bottom row). A low detection limit of 1.2 nM SWNT-RGD was reached for
labeled cells on quartz substrate, below which cells (visible in optical images in insets
of Figure 4h, 4j) were hardly discerned in the SWNT fluorescence images. In contrast,
on Au/Au substrates, SWNT fluorescent cell images were observed for cells treated
by SWNTs down to ~48 pM in concentration, suggesting a 25-fold improvement in
the detection limit on Au/Au film (Figure 4k).
240 pM
4x
i)
2x
g)
48 pM
8x
On Au/Au
On quartz
a)
k)k)
b)
k)
30 nM
6 nM
1.2 nM
30µm
c)
e)
4x
d)
4x
f)
8x
h)
16x
j)
32x
Figure 4. SWNT NIR fluorescence images of αvβ3-integrin positive U87-MG cells,
on Au/Au (top panel) and quartz (bottom panel), stained with different concentrations
of SWNT-RGD conjugates: a)&b) 30 nM, c)&d) 6 nM, e)&f) 1.2 nM, g)&h) 240 pM
and i)&j) 48 pM. Some of these images with weak fluorescence were autoscaled by a
certain multiplier as labeled. Insets in h) and j) showed the corresponding optical
images in the same field of view. All fluorescence images were false-colored. k) Cell
titration curves showed the average fluorescence intensity of SWNTs in each stained
U87-MG cell on Au/Au film (red curve) and on quartz (blue curve) at different
staining concentrations.
11
In conclusion, we employed plasmonic gold substrates for the first time to perform
near-IR fluorescence enhanced molecular imaging of cells in the 0.8-1.4 µm spectral
window based on carbon nanotubes and organic fluorophores. The novel
solution-grown gold substrate was general in enhancing both carbon nanotubes and
infrared dye IR800, by ~9 times and ~6 times respectively, affording higher
sensitivity and specificity of molecular cell imaging in the advantageous 0.8-1.4 µm
spectral window. Cell labeling at different incubation temperatures blocked or
allowed endocytosis of nanotube fluorophores, leading to observation of a distance
dependent fluorescence enhancement inside cells. This effect could be used to observe
transmembrane behavior of single NIR fluorophores in live cells when the
fluorescence enhancement decay distance matches cell membrane thickness. Further
possibilities with NIR-FE imaging include single molecule imaging and tracking of
SWNTs or other NIR dyes on cell membrane, molecular imaging of low abundance
cell membrane proteins, and even some in vivo NIR-FE imaging using smaller yet
enhancing Au nanostructures as enhancing platform to serve in a fluidic system.
12
[6]
[9]
[7]
[8]
References
J. V. Frangioni, Curr. Opin. Chem. Biol. 2003, 7, 626.
[1]
V. Ntziachristos, J. Ripoll, R. Weissleder, Opt. Lett. 2002, 27, 1652.
[2]
J. E . Aubin, J. Histochem. Cytochem. 1979, 27, 36.
[3]
Z. Liu, S. Tabakman, K. Welsher, H. J. Dai, Nano Res. 2009, 2, 85.
[4]
Z. Liu, S. Tabakman, S. Sherlock, X. L. Li, Z. Chen, K. L. Jiang, S. S. Fan, H.
[5]
J. Dai, Nano Res. 2010, 3, 222.
J. H. Flanagan, S. H. Khan, S. Menchen, S. A. Soper, R. P. Hammer,
Bioconjugate Chem. 1997, 8, 751.
Y. H. Lin, R. Weissleder, C. H. Tung, Bioconjugate Chem. 2002, 13, 605.
R. Weissleder, C. H. Tung, U. Mahmood, A. Bogdanov, Nat. Biotechnol. 1999,
17, 375.
V. Ntziachristos, C.-H. Tung, C. Bremer, R. Weissleder, Nat. Med. 2002, 8,
757.
[10] M. Bruchez, M. Moronne, P. Gin, S. Weiss, A. P. Alivisatos, Science 1998, 281,
2013.
[11] W. C. W. Chan, S. M. Nie, Science 1998, 281, 2016.
[12] X. Y. Wu, H. J. Liu, J. Q. Liu, K. N. Haley, J. A. Treadway, J. P. Larson, N. F.
Ge, F. Peale, M. P. Bruchez, Nat. Biotechnol. 2003, 21, 41.
[13] M. J. O'Connell, S. M. Bachilo, C. B. Huffman, V. C. Moore, M. S. Strano, E.
H. Haroz, K. L. Rialon, P. J. Boul, W. H. Noon, C. Kittrell, J. P. Ma, R. H.
Hauge, R. B. Weisman, R. E. Smalley, Science 2002, 297, 593.
[14] S. M. Bachilo, M. S. Strano, C. Kittrell, R. H. Hauge, R. E. Smalley, R. B.
Weisman, Science 2002, 298, 2361.
[15] P. Cherukuri, S. M. Bachilo, S. H. Litovsky, R. B. Weisman, J. Am. Chem. Soc.
2004, 126, 15638.
[16] K. Welsher, Z. Liu, D. Daranciang, H. Dai, Nano Lett. 2008, 8, 586.
[17] P. Cherukuri, C. J. Gannon, T. K. Leeuw, H. K. Schmidt, R. E. Smalley, S. A.
Curley, R. B. Weisman, P. Natl. Acad. Sci. USA 2006, 103, 18882.
[18] T. K. Leeuw, R. M. Reith, R. A. Simonette, M. E. Harden, P. Cherukuri, D. A.
Tsyboulski, K. M. Beckingham, R. B. Weisman, Nano Lett. 2007, 7, 2650.
[19] K. Welsher, Z. Liu, S. P. Sherlock, J. T. Robinson, Z. Chen, D. Daranciang, H.
J. Dai, Nat. Nanotechnol. 2009, 4, 773.
J. T. Robinson, K. Welsher, S. M. Tabakman, S. P. Sherlock, H. L. Wang, R.
Luong, H. J. Dai, Nano Res. 2010, 3, 779.
[21] X. X. Peng, H. X. Chen, D. R. Draney, W. Volcheck, A. Schutz-Geschwender,
D. M. Olive, Anal. Biochem. 2009, 388, 220.
[22] R. Philip, A. Penzkofer, W. Baumler, R. M. Szeimies, C. Abels, J. Photoch.
Photobio. A 1996, 96, 137.
[23] R. B. Mujumdar, L. A. Ernst, S. R. Mujumdar, C. J. Lewis, A. S. Waggoner,
Bioconjugate Chem. 1993, 4, 105.
[24] D. Magde, R. Wong, P. G. Seybold, Photochem. Photobiol. 2002, 75, 327.
J. Crochet, M. Clemens, T. Hertel, J. Am. Chem. Soc. 2007, 129, 8058.
[25]
[26] L. J. Carlson, S. E. Maccagnano, M. Zheng, J. Silcox, T. D. Krauss, Nano Lett.
[20]
13
[36]
2007, 7, 3698.
[27] V. Perebeinos, J. Tersoff, P. Avouris, Nano Lett. 2005, 5, 2495.
[28] M. S. Strano, C. B. Huffman, V. C. Moore, M. J. O'Connell, E. H. Haroz, J.
Hubbard, M. Miller, K. Rialon, C. Kittrell, S. Ramesh, R. H. Hauge, R. E.
Smalley, J. Phys. Chem. B 2003, 107, 6979.
[29] S. Y. Ju, W. P. Kopcha, F. Papadimitrakopoulos, Science 2009, 323, 1319.
[30] G. S. Hong, S. M. Tabakman, K. Welsher, H. L. Wang, X. R. Wang, H. J. Dai,
J. Am. Chem. Soc. 2010, 132, 15920.
[31] E. Le Moal, E. Fort, S. Leveque-Fort, F. P. Cordelieres, M. P. Fontaine-Aupart,
C. Ricolleau, Biophys. J. 2007, 92, 2150.
J. Zhang, Y. Fu, D. Liang, R. Y. Zhao, J. R. Lakowicz, Langmuir 2008, 24,
12452.
[33] B. Radha, M. Arif, R. Datta, T. K. Kundu, G. U. Kulkarni, Nano Res. 2010, 3,
738.
[34] Z. Liu, W. B. Cai, L. N. He, N. Nakayama, K. Chen, X. M. Sun, X. Y. Chen, H.
J. Dai, Nat. Nanotechnol. 2007, 2, 47.
[35] S. M. Tabakman, Z. Chen, H. S. Casalongue, H. Wang, H. Dai, Small 2011, 7,
499.
J. H. Kim, J. H. Ahn, P. W. Barone, H. Jin, J. Q. Zhang, D. A. Heller, M. S.
Strano, Angew. Chem. Int. Edit. 2010, 49, 1456.
[37] N. W. S. Kam, T. C. Jessop, P. A. Wender, H. J. Dai, J. Am. Chem. Soc. 2004,
126, 6850.
[38] N. W. S. Kam, H. J. Dai, J. Am. Chem. Soc. 2005, 127, 6021.
[39] N. W. S. Kam, Z. Liu, H. J. Dai, Angew. Chem. Int. Edit. 2006, 45, 577.
[40] H. Jin, D. A. Heller, M. S. Strano, Nano Lett. 2008, 8, 1577.
[41] H. Jin, D. A. Heller, R. Sharma, M. S. Strano, Acs Nano 2009, 3, 149.
[42] H. Mertens, A. F. Koenderink, A. Polman, Phys. Rev. B 2007, 76.
J. Gersten, A. Nitzan, J. Chem. Phys. 1981, 75, 1139.
[43]
[44] A. Chen, V. T. Moy, Biophys. J. 2000, 78, 2814.
[45] Y. Fang, N. H. Seong, D. D. Dlott, Science 2008, 321, 388.
[46] Z. Chen, S. M. Tabakman, A. P. Goodwin, M. G. Kattah, D. Daranciang, X. R.
Wang, G. Y. Zhang, X. L. Li, Z. Liu, P. J. Utz, K. L. Jiang, S. S. Fan, H. J. Dai,
Nat. Biotechnol. 2008, 26, 1285.
[47] K. E. Adams, S. Ke, S. Kwon, F. Liang, Z. Fan, Y. Lu, K. Hirschi, M. E.
Mawad, M. A. Barry, E. M. Sevick-Muraca, J. Biomed. Opt. 2007, 12.
[48] Y. Chen, S. Dhara, S. R. Banerjee, Y. Byun, M. Pullambhatla, R. C. Mease, M.
G. Pomper, Biochem. Bioph. Res. Co. 2009, 390, 624.
[32]
|
1904.08216 | 1 | 1904 | 2019-04-17T12:09:53 | Exactly solvable model of gene expression in proliferating bacterial cell population with stochastic protein bursts and protein partitioning | [
"physics.bio-ph",
"q-bio.MN"
] | Many of the existing stochastic models of gene expression contain the first-order decay reaction term that may describe active protein degradation or dilution. If the model variable is interpreted as the molecule number, and not concentration, the decay term may also approximate the loss of protein molecules due to cell division as a continuous degradation process. The seminal model of that kind leads to gamma distributions of protein levels, whose parameters are defined by the mean frequency of protein bursts and mean burst size. However, such models (whether interpreted in terms of molecule numbers or concentrations) do not correctly account for the noise due to protein partitioning between daughter cells.
We present an exactly solvable stochastic model of gene expression in cells dividing at random times, where we assume description in terms of molecule numbers with a constant mean protein burst size. The model is based on a population balance equation supplemented with protein production in random bursts.
If protein molecules are partitioned equally between daughter cells, we obtain at steady state the analytical expressions for probability distributions similar in shape to gamma distribution, yet with quite different values of mean burst size and mean burst frequency than would result from fitting of the classical continuous-decay model to these distributions. For random partitioning of protein molecules between daughter cells, we obtain the moment equations for the protein number distribution and thus the analytical formulas for the squared coefficient of variation. | physics.bio-ph | physics |
Exactly solvable model of gene expression in proliferating bacterial cell population
with stochastic protein bursts and protein partitioning
Jakub Jędrak,1, ∗ Maciej Kwiatkowski,2 and Anna Ochab-Marcinek1
1Institute of Physical Chemistry, Polish Academy of Sciences, Kasprzaka 44/52, 01-224 Warsaw, Poland
2unaffiliated; ORCID: 0000-0001-7675-4595
(Dated: April 18, 2019)
Many of the existing stochastic models of gene expression contain the first-order decay reaction
term that may describe active protein degradation or dilution. If the model variable is interpreted
as the molecule number, and not concentration, the decay term may also approximate the loss of
protein molecules due to cell division as a continuous degradation process. The seminal model of
that kind leads to gamma distributions of protein levels, whose parameters are defined by the mean
frequency of protein bursts and mean burst size. However, such models (whether interpreted in
terms of molecule numbers or concentrations) do not correctly account for the noise due to protein
partitioning between daughter cells.
We present an exactly solvable stochastic model of gene expression in cells dividing at random
times, where we assume description in terms of molecule numbers with a constant mean protein burst
size. The model is based on a population balance equation supplemented with protein production
in random bursts.
If protein molecules are partitioned equally between daughter cells, we obtain at steady state
the analytical expressions for probability distributions similar in shape to gamma distribution, yet
with quite different values of mean burst size and mean burst frequency than would result from
fitting of the classical continuous-decay model to these distributions. For random partitioning of
protein molecules between daughter cells, we obtain the moment equations for the protein number
distribution and thus the analytical formulas for the squared coefficient of variation.
I.
INTRODUCTION
Genetically identical cells in a proliferating microbial
population may differ not only in their cell-cycle age,
volume or growth rate, but also in the copy numbers of
molecules, in particular, the protein molecules. Even if
the dependence of gene expression on growth rate [1 --
4] and cell-cycle stage [5] can be neglected, there still
remain at least two factors causing fluctuations in the
molecule numbers between individual cells: First, gene
expression is an inherently stochastic process because of
the small number of molecules involved in biochemical
reactions. Second, molecules, including proteins, are par-
titioned between daughter cells at cell division in a more
or less random manner [6, 7].
Yet, many stochastic models of gene expression neglect
the protein partitioning [8]. The classical models with a
first-order decay reaction term [9 -- 12] may be interpreted
in two ways: (i) If the model describes the protein levels
in terms of molecule numbers, then the degradation term
describes not only a possible active protein degradation
(rare in bacteria [13]) but it also mimicks the effect of cell
division. However, this description neglects the fluctua-
tions both due to the abrupt protein loss at cell division
(which are present even at the half-by-half protein par-
titioning) and due to a possible random partitioning of
the molecules between cells.
(ii) If the same classical
∗[email protected]
model describes the protein levels in terms of concentra-
tions, then the continuous degradation term reflects, be-
sides possible active degradation, the dilution due to cell
growth. Under the assumption of a perfectly equal pro-
tein partitioning, the concentrations remain unchanged
at cell division. However, if protein partitioning is ran-
dom, then it should result in concentration fluctuations
(at binomial partitioning, they extinct at high gene ex-
pression levels), and these fluctuations are neglected in
the conventional model.
Based on the classical models of gene expression where
proteins are assumed to be produced in random transla-
tional bursts [9 -- 12] and on the population balance equa-
tions [14 -- 16], we propose a minimal analytically tractable
model of gene expression in a population of dividing cells.
This model directly accounts for the protein loss at cell
division and random partitioning of the molecules be-
tween daughter cells.
The manuscript is organized as follows: Sec.
II
presents the theoretical model. In Sec. III A we calculate
the moments of the steady-state protein number distribu-
tion for any symmetric random partitioning of proteins
between daughter cells.
III B we show that a
noise floor, i.e. the absolute lower bound for noise as a
function of mean protein number, occuring for highly ex-
pressed proteins, arises in the model as a consequence of
protein loss at cell division. This kind of a noise floor is a
property of the description in terms of molecule numbers.
In Sec. III C we compare the protein number distribu-
tions resulting from our model to the distributions from
the seminal Friedman's model [11].
III C 1 we
In Sec.
In Sec.
show that if both distributions have the same mean and
variance then the underlying mean burst frequency and
mean burst size may differ. In Sec. III C 2 we obtain the
analytical form of the protein number distribution for
the half-by-half protein partitioning and we show that
its shape is very similar to the shape of a gamma dis-
tribution from the Friedman's model [11]. In Sec. III D
we analyze the behavior of the distribution tails for large
protein numbers. In Sec. III E we compare the squared
coefficient of variation of our model to the existing exper-
imental data. Section IV contains discussion and conclu-
sions.
II. MODEL
We combine the analytical framework first introduced
by Friedman et al. [11] with the population balance equa-
tion (PBE) [14 -- 16], a standard tool for modelling micro-
bial populations.
In Ref. [11], gene expression was described as a com-
pound Poisson process with random Poissonian events of
protein production, where a random number of protein
molecules was produced at once in each such event. The
protein burst size u was drawn from a distribution ν(u).
Originally, ν(u) was assumed to be an exponential dis-
tribution [11], based on the experimental data for E. coli
[10, 17], but here we will relax this assumption. Protein
production was counterbalanced in the Friedman's model
[11] by a continuous, deterministic protein decay term.
The model allowed self-regulation: The burst occurrence
rate k was modulated by a transfer function hp(x), de-
pendent on the current protein level x. We note that the
continuous variable x was assumed to be protein concen-
tration in Ref. [11]. However, in this study, we interpret
x as the molecule number. This distinction is crucial be-
cause cell division exactly by half maintains a constant
protein concentration, whereas the protein number drops
by 1/2. If protein production bursts are constant in terms
of molecule number, then their size in the units of con-
centration decreases in growing cells.
Therefore, we treat here the theoretical framework of
Ref. [11] as a continuous approximation to discrete fixed
cell size gene expression models, cf. e.g.
[18], and not
in its original interpretation, as a model describing evo-
lution of protein concentration in growing and dividing
cell. Within the model proposed in Ref.
[11], the first
order decay term describes both true protein degradation
(if present) and protein dilution due to cell growth.
On the other hand, population balance equation (PBE)
describes the time evolution of the cell number density
in a proliferating microbial population [14, 15]. At cell
division, protein molecules are partitioned according to
x → {qx, (1 − q)x}, where x ∈ [0, ∞) is a continu-
ous protein copy number, and not protein concentration,
whereas the division ratio 0 ≤ q ≤ 1 is a random number,
drawn from the division ratio probability density func-
tion (PDF) η(q) = η(1 − q). Within the PBE approach
2
as used in [14 -- 16], protein production is described as a
deterministic process. However, both protein production
rate and cell division rate may depend on x in a nontrvial
way.
By merging both the above theoretical frameworks, we
get the time-evolution equation for p(x, t), a PDF of the
protein number x at time t in the population of cells
(Appendix A):
∂p(x, t)
∂t
= −
∂
∂x
[g(x)p(x, t)] − hd(x)p(x, t)
hd(cid:18) x
q(cid:19) p(cid:18) x
q
, t(cid:19) dq
hd(ξ)p(ξ, t)dξ
0
q
η(q)
+ 2Z 1
− p(x, t)Z ∞
+ kZ x
0
0
w(x − x′)hp (x′) p(x′, t)dx′,
(1)
In the above equation, x − x′ is the burst size, ν(x − x′)
is the burst size PDF, w(x − x′) ≡ ν(x − x′) − δ(x − x′)
and δ(x − x′) is Dirac delta distribution [11]. The x-
dependence of cell division and protein production rates
are given by hd(x) and hp(x), respectively. Thus, khp(x)
is the burst frequency, whereas g(x) describes either de-
terministic protein production or degradation. Eq. (1) is
a generalization of both Eq. (1) of Ref. [11] and Eq. (1)
of Ref. [16].
If x cannot be treated as proportional to the cell mass
or volume, it is unclear how the division rate should de-
pend on the copy number of a given protein, and various
functional forms of hd(x) were used in the literature [14 --
16]. For that reason, and in order to obtain an analyti-
cally tractable model, we focus here on the simplest case
of the constant cell division rate, 2hd(x) ≡ ∆ = const,
which implies R ∞
0 hd(ξ)p(ξ)dξ = ∆/2.
III. RESULTS
A. Moments of the steady-state protein number
distribution
We consider the steady-state solution of Eq.
(1),
∂p(x, t)/∂t = 0. Assuming g(x) = −γx, γ ≥ 0 and
unregulated protein production (hp(x) = 1), we have
−γ
d
dx
[xp(x)] = ∆Z 1
+ kZ x
0
0
q(cid:19) − p(x)(cid:21) dq
p(cid:18) x
η(q)(cid:20) 1
w(x − x′)p(x′)dx′.
q
(2)
The Laplace transform (L{. . .}) of Eq. (2) yields
γs
dG(s)
ds
= ∆Z 1
0
η(q) [G(qs) − G(s)] dq + k w(s)G(s),
(3)
where G(s) = L{p(x)}, and w(s) = L{w(u)} = ν(s) − 1.
Moment equations follow from Eq. (3),
[(1 − Mr)∆ + rγ] µr = k
r
Xl=1
(cid:18)r
l(cid:19)µr−lml,
(4)
where µr = R ∞
tein number PDF, mr = R ∞
of the burst size PDF, and Mr = R 1
0 xrp(x)dx is the r-th moment of the pro-
0 urν(u)du is the r-th moment
0 qrη(q)dq is the r-th
moment of the division ratio PDF (note that M1 = 1/2).
For γ 6= 0, ∆ = 0 (the proteins are degraded, but there
is no protein partitioning at cell division) Eq. (4) yields
µ1 = am1 ≡ µ1,γ,
κ2 = µ2 − µ2
1 =
1
2
am2 ≡ κ2,γ,
(5)
where a = k/γ [19]. On the other hand, for γ = 0 (the
proteins are stable) and ∆ 6= 0, from Eq. (4) we get
µ1 = 2ωm1 ≡ µ1,∆,
κ2 =
4ω2M2m2
1 + ωm2
(1 − M2)
≡ κ2,∆,
(6)
where ω = k/∆. The variance of the protein number,
κ2,∆ = var[p(x)] (6) is a monotonically increasing func-
1)/M2
tion of (M2 − M2
1 = 4M2 − 1, i.e., of the noise due
to protein partitioning.
For a given ν(u) and η(q), if a = 2ω, i.e., if µ1,γ (5) is
equal to µ1,∆ (6), then from Eqs. (5) and (6) it follows
that κ2,∆ > κ2,γ. Therefore, within the present descrip-
tion in terms of molecule numbers, the first-order decay
such as in the classical model by Friedman et al. [11] is
not equivalent to any protein partitioning mechanism.
B. Squared coefficient of variation has a noise floor
due to cell division
Protein degradation is important for eukaryotic organ-
isms, but may be neglected in the case of bacteria; hence
we put γ = 0. From Eq.
(6) we obtain the squared
coefficient of variation
c2
v(µ1) ≡
κ2,∆
µ2
1,∆
=
m2
2m1
1
(1 − M2)
1
µ1
+
M2
(1 − M2)
.
(7)
The protein number noise, measured by the squared co-
efficient of variation (7) is thus of the form
c2
v(µ1) =
E1
µ1
+ E0,
(8)
with µ1-independent E0 and E1, which leads to the char-
acteristic boomerang-like shape of the log-log plot (Fig.
1A). In Fig. 1, the half-by-half protein partitioning is
assumed,
η(q) = ηd(q) ≡ δ (q − 1/2) ,
(9)
which does not depend on x and which gives M2 = 1/4.
However, for small x, the half-by-half division as given by
b=10
b=18.0 +/- 2.1
A
β
104
103
102
101
100
b=10
b=5
b=1
ω
B
3
104
103
102
101
100
102
)
12
µ
/
2
b=5
101
b=1
100
i
κ
(
e
s
o
n
e
c
n
e
c
s
e
r
o
u
F
l
10-1
10-1 100 101 102 103 104
Mean fluorescence (µ
1)
10-1
10-1
100
α
10-1
101
FIG. 1: (A) Plot of c2
v(µ1) (7) as a function of µ1 = 2ωm1 for
m2/m1 = 2b and for selected mean burst sizes b = m1. Half-
by-half protein partitioning is assumed, M2 = 1/4, which
does not depend on x. Dashed line: E0 = E0,d = 1/3, the
lowest value of the noise floor allowed by the model. Points:
[21], see Sec. III E). Symbols
Experimental data from Ref.
and colors in (A) are used as in Ref.
[21]. Symbols, carbon
source: arabinose ( ), maltose ((cid:4)), glycerol ((cid:7)), glucose (N),
glucose+malate (H). Colors, promoter activity induced by:
10 (blue), 20 (gold), 50 (green), 100 (violet), 1000 (turquoise)
µM IPTG. Gray dashed curve: Fit of our model to the data.
(B) Mapping of mean burst frequency ω and mean burst size
b of our model (where cell division is explicitly taken into
account) onto the corresponding values α and β of these pa-
rameters appearing in Eq.
(12) (where there is no protein
loss at cell division but a continuous protein degradation is
assumed). Curves: β(α) (14) for variable ω and constant b.
Colors denote the value of ω. Circles, squares, and triangles
denote ω = 10i, i = −1, 0, 1, 2, 3, 4. Dashed line: upper bound
for α within our model, αmax = E −1
0,d = 3. Black dots: The
same data as in (A) but the colors and symbols were omit-
ted for the sake of clarity. Gray dashed curve: Fit from (A),
mapped using Eq. (14).
(9) is much less realistic than other continuous approxi-
mations to discrete binomial distribution, such as the 'all
or none' partitioning (η(q) = ηb(q) ≡ 1
2 [δ(q) + δ(1 − q)])
or even a uniform distribution of partition ratio (η(q) =
ηu(q) ≡ Θ(q)Θ (1 − q)), considered in Refs. [16, 20].
E0 = M2/(1 − M2) provides the lower bound for ex-
trinsic noise and does not depend on the details of pro-
tein production mechanism. However, the contribution
of protein partitioning to the value of E0 predicted by the
present model for a realistic level of protein partitioning
noise is likely to be overestimated for the following rea-
son: Eq. (2) is a Master equation with time-independent
parameters, hence it describes the cell division events as a
Poisson process [22]. In consequence, the cell cycle length
has an unrealistic, exponential distribution, contributing
with too much noise (see Appendix C). Therefore, the
age structure of the population is also unrealistic (Ap-
pendix D). This indicates a limitation of the simplest,
single-variable version of the PBE approach as a candi-
date for a minimal model of gene expression as it assumes
the "most random" distribution of cell cycle lengths (i.e.,
the one that naturally occurs from a Poisson process).
Thus, our model provides an "upper bound" for the
noise contribution due to cell cycle length fluctuations
(meaning the "most random" fluctuations that can be
generated by a model based on Master equation [23]):
(9) there is no
In the case when η(q) is given by Eq.
protein partitioning noise (var[δ(cid:0)q − 1
has the minimal value E0 = 1/3 ≡ E0,d. Therefore,
E0,d is the contribution to the noise floor coming solely
from fluctuations of the cell cycle length, whereas the
remaining part, E0 − E0,d = (4M2 − 1)/[3(1 − M2)] is
due to protein partitioning noise.
2(cid:1)] = 0) and E0
C. Our model vs. Friedman's model [11]: Similar
distributions but different parameters
1. Mean burst frequency and mean burst size differ between
the models despite the same mean and variance of protein
number distribution
Let us consider a single-parameter burst size PDF,
ν(u; b) = h (u/b) /b, where u ≡ x − x′ is a burst size,
such that m1 = b and m2 = Cb2. C = 2 for the expo-
nential burst distribution:
ν(u) = b−1e−u/b,
ν(s) ≡ L[ν(u)] = (sb + 1)−1.
(10)
For m2/m1 = Cb = const, the increase in mean protein
number µ1 = 2ωb is caused solely by the increase in ω
(i. e., by the increase in burst frequency k or decrease in
cell division rate ∆). In such a case, we move along the
boomerang-shaped trajectories for a fixed value of b, like
those depicted in Fig. 1A. In general, µ1 may be varied
both due to the change in b and ω.
If k = 0 and g(x) = −γx term in Eq. (2) is replaced
with g(x) = σ > 0 (this is a special case of the model of
Ref. [16]), then instead of (7) we obtain
c2
v(µ1) = E0 =
M2
1 − M2
= const.
(11)
In other words, the squared coefficient of variation is then
described by some general equation in the form of (8)
(but not (7)), where E0 6= 0, E1 = 0 if protein production
in our model is deterministic and cell division is stochas-
tic. The opposite situation, E1 6= 0, E0 = 0, is encoun-
tered for γ 6= 0, ∆ = 0, in particular for the model of
unregulated, bursty gene expression with continuous pro-
tein degradation [11], which predicts that protein num-
bers are gamma-distributed:
pγ(x; α, β) ≡
xα−1 exp (−x/β)
βαΓ(α)
.
(12)
Then, c2
v(µ1) = 1/α = β/µ1, i. e., E1 = β and E0 = 0.
To make a comparison of pγ(x; α, β) (12) to the protein
number PDF of the present model, p(x), we assume now
4
that these two distributions have equal means and equal
second moments, µr,γ = µr,∆, r = 1, 2. We also assume
that the burst size PDF ν(u) is given by (10), but the
mean burst size may be different for the two models,
m1,γ = β 6= b = m1,∆. These assumptions yield the
following mapping:
α =
µ2
1
κ2
=
ω (1 − M2)
ωM2 + 1/2
,
β =
κ2
µ1
=
(2M2ω + 1) b
1 − M2
.
From (13) we obtain
β(α) =
b
1 − (1 + α)M2
.
(13)
(14)
For simplicity, we assume now that protein partitioning
is deterministic, i.e., that each daughter cell obtains ex-
actly half of the total number of protein molecules at cell
division. In such a case, η(q) = ηd(q) is given by Eq. (9),
so that M2 = 1/4. In Fig. 1B we plot α vs. β (14) to
show that the mean burst frequency ω and mean burst
size b in our model take different values than the corre-
sponding parameters α and β of the gamma distribution
(12). In particular, ω is not limited to several bursts per
cell cycle as was α, so the mean burst size b does not
need to take as high values as would β need to take in
order to obtain a high level of gene expression.
2. Analytical form of the protein number distribution for
half-by-half protein partitioning. Apparent similarity to
gamma distribution.
Below, we will show that equating the two first mo-
ments of our PDF p(x) with those of the gamma dis-
tribution pγ(x) (12) yields similar overall shapes of the
distributions. If so, any experimental gamma-shaped dis-
tributions with c2
v ≥ E0,d = 1/3 may be as well fitted
by the distributions given by our model (under the as-
sumption that protein numbers, and not concentrations,
were measured in experiment). With η(q) = δ(q − 1
2 ),
we can rewrite Eq.
(3) as G(s) = R(s)G (s/2), where
R(s) ≡ {1 − ω[ν(s) − 1]}−1, R(0) = 1. Solving by itera-
i=0 R(cid:0)s/2i(cid:1). For ν(s) (10), G(s)
tion, we get G(s) = Q∞
reads
G(s) =
=
∞
∞
Yi=0
Xr=0
2−ibs + 1
2−i(1 + ω)bs + 1
=
(−bs; 1
2 )∞
(−b(1 + ω)s; 1
2 )∞
((1 + ω)−1; 1
2 )r
2 ; 1
( 1
2 )r
(−b(1 + ω)s)r,
(15)
where (a; q)k is a q-Pochhammer symbol. In (15) we used
the q-binomial theorem [24],
(az; q)∞
(z; q)∞
=
∞
Xn=0
(a; q)n
(q; q)n
zn.
(16)
The symbol q appering in Eq. (16) should not be con-
fused with protein partitioning ratio q appearing e.g. in
Eqs.
(1), (2), (3) or (9). The letter q is traditionally
used in the branch of mathematics called q-theory or q-
analogs (q-binomial theorem or q-Pochhammer symbol
are examples of such q-analogs).
From Eq. (15) we obtain cumulants of L−1{G(s)},
κr =
2rbr(r − 1)![(1 + ω)r − 1]
(2r − 1)
.
(17)
[16] and [20], instead of (18) we obtain a
used in Refs.
statistical mixture of two broad (c2
v ≥ 1/2) gamma dis-
tributions (Appendix G). Therefore in the present case it
is no longer true that the "steady-state population distri-
bution (...) becomes insensitive to the division details"
[20]. For large x, η(q) = Θ(q)Θ (1 − q) is much less real-
istic than η(q) = ηd(q) = δ(q − 1
2 ) (9) used here, and it
leads to a higher noise floor (E0 = 1/2).
5
We also have
L−1{G(s)} = p(x) =
∞
Xi=0
Ci(ω)
b
exp(cid:18) −2ix
b(ω + 1)(cid:19) ,
(18)
where
Ci(ω) =
2iω
(2(1 + ω)−1; 2)i
( 1
2 (1 + ω)−1; 1
2 )∞
(1 + ω)2
(2; 2)i
2 ; 1
( 1
2 )∞
.
(19)
For ω = ωn ≡ 2n − 1, n = 1, 2, 3, . . ., p(x) (18) can be
written as a finite series
p(x) ≡ pn(x) =
1
b
n
Xl=1
An,l exp(cid:16)−
x
2lb(cid:17) ,
(20)
where
An,l =
(−1)n−l2
l(l−3)
2
Ql−1
i=1(2i − 1)Qn−l
j=1(2j − 1)
.
(21)
Equations (18) and (19) define a two-parameter family of
PDFs, resembling the gamma PDF (Fig. 2). In partic-
ular, p(x) (18) is right-skewed, unimodal for ω > 1, and
monotonically decreasing for ω ≤ 1.
D. Distribution tail for large protein numbers
In both p(x) (18) and pγ(x) (12), the mean burst size
(b and β, respectively) is the scaling parameter. The tail
of p(x) (18) is exponential: For large x we have p(x) ∼
exp[−x/(b(ω + 1))]. However, in contrast to the gamma
distribution, where the leading term is exp[−x/b], the
exponent in our model depends not only on the mean
burst size, but also on burst frequency.
The same asymptotic behaviour as for p(x) (18) is
present if instead of η(q) = δ(q − 1
2 ) any other η(q)
for
is used (except for some pathological cases, e.g.
ηb(q) = 1
2 [δ(q) + δ(1 − q)]) (Appendix F). This is the
special case of a yet more general result. Using Eq. (1), it
can be shown in a similar manner as in Ref. [16] that for
g(x) = 0 and ν(u) (10) the ratio p(x2)/p(x1) at the tail
of the PDF is well approximated by exp (−I12), where
I12 = Z x2
and R = R ∞
x1
kh′
p(x) + h′
d(x) + hd(x)/b + R/b
khp(x) + hd(x) + R
dx,
(22)
0 hd(ξ)p(ξ)dξ (Appendix F). However, even if
the tails are universal, i.e., independent on the division-
ratio PDF η(q), the corresponding probability distribu-
tions are not. For example, for η(q) = Θ(q)Θ (1 − q)
E. Comparison to experimental data
In this section, we compare the squared coefficient of
variation as a function of the mean protein level in our
model to the existing experimental data, under the as-
sumption of equal protein partitioning between daugh-
ter cells (the lowest possible noise due to cell division,
Fig. 1A).
In Ref.
[21], the authors measured total cell fluores-
cence emitted by the green fluorescent protein (GFP),
encoded for by a gene controlled by the hyper-spank pro-
moter in B. subtilis. The promoter activity was modu-
lated by the concentration of isopropyl-D-thiogalactoside
(IPTG). An independent parameter varied in the exper-
iment was the cell growth rate, depending on the carbon
source in the medium. The data points shown in Fig. 1A
were obtained in Ref. [21] by variation of these two pa-
rameters (we use the symbols and colors as in that Ref-
erence; we manually extracted the data points from Fig.
4a therein using the xyscan software).
Assuming that the total cell fluorescence scales linearly
with the number of reporter proteins, the x variable in
our model may be reinterpreted as the fluorescence level
and the mean protein burst size m1 ≡ b is then expressed
in fluorescence units. Thus, the mean protein number is
proportional to the mean total cell fluorescence, and the
squared coefficient of variation does not depend on the
units in which gene expression was measured (molecule
number or fluorescence units). We note that the lin-
ear scaling between the GFP molecule number and total
cell fluorescence is not necessarily an obvious assumption
(see, e.g., Ref. [25]). On the other hand, the authors of
Ref.
[21] estimated that GFP maturation time did not
affect strongly the fluorescence level, which may possibly
exclude one source of non-linear scaling of the fluores-
cence detected vs. molecule number. The linear scaling
was also assumed in Ref. [26].
The range of the mean total cell fluorescence mea-
sured in Ref.
[21] was too narrow to show a noise floor
(Fig. 1A). However, the data provide a hint that if the
noise floor exists, it would be lower than that predicted by
our model. This suggests that the distribution of cell cy-
cle lengths in our model is indeed too wide, as we pointed
out in Section III B.
For a rough comparison, we also plotted a fit of the
squared coefficient of variation c2
v of our model (7) to
the data. The slope of c2
v in our model seems to be in
agreement with the experimental results. The fit yields
an estimation of the mean burst size m1 ≡ b = 18.0 ± 2.1
in the units of fluorescence, as used in Ref.
[21]; how-
ever, this value should be treated with caution because
some other model with a lower noise floor due to a nar-
rower distribution of cell-cycle lengths may possibly yield
a different value of b.
For completeness, we also plotted the data from Ref.
[21] in Fig. 1B (we omitted the colors and symbols used
in Fig. 1A). We made the assumption that the total cell
fluorescence in the experiment was gamma-distributed,
as shown in the SI of Ref. [21]. We mapped the mean hf i
v,f = var(f )/hf i2 of
and squared coefficient of variation c2
the total cell fluorescence onto the corresponding values
v,f )−1, β =
of parameters of gamma distributions: α = (c2
hf i · c2
v,f .
A noise floor was observed in other experiments re-
ported in literature [3, 12, 25 -- 29] but most of these data
were not suitable for comparison to our model (see an
overview in Appendix E) because they were normalized
to cell volume or gated to make them independent on
cell-cycle stage. Thus, molecule number fluctuations due
to loss of proteins at cell division were absent in these
data.
There are two studies on S. cerevisiae that reported
the gene expression noise floor in ungated measurements.
Its levels lower than in our model (where E0 = 1/3): In
Ref. [27], Fig. 2 and S2 therein, the noise floor measured
as the coefficient of variation cv took the values between
0.3 and 0.4, corresponding to the squared coefficient of
variation c2
v between 0.09 and 0.16. In Ref. [3], Fig. 2A
and S9 therein, c2
v was between 0.1 and 0.2. One reason
for that difference may be the too wide distribution of cell
cycle lengths in our model. But it should also be noted
that our model may be far too simplistic for eukaryotic
cells, as it does not account for the discrete promoter
activity states due to chromatin remodeling, the nuclear
transport, etc.
IV. DISCUSSION AND CONCLUSIONS
[16].
In Ref.
[11] and Ref.
We have proposed a model of gene expression in a pro-
liferating cell population, which is a generalization of the
models of Refs.
[11], the
protein production was stochastic, whereas the decrease
of protein concentration was due to a deterministic pro-
tein degradation process. Here, we treat the model of
Ref. [11] as a continuous approximation (still in the units
of protein number) to discrete gene expression models,
such as in Ref. [18], that describe non-growing and non-
dividing cells.
In Ref.
[16], the situation was just opposite: Protein
partitioning was stochastic and protein production was
deterministic. After combining the stochastic protein
production (of Ref. [11]) with the stochastic protein par-
titioning (used in Ref.
[16]), we obtain the boomerang-
like shape of the log-log plot of mean protein copy number
vs. protein copy number noise.
6
n=1
n=2
n=3
n
ω
α
β/b
1
1
1
2
3
3
7
4
15
1.8
2.333
2.647
2
3.333
6
11.333
20 × 103
)
β
,
α
;
x
(
γ
p
,
)
b
,
)
n
(
ω
;
x
(
n
p
15
10
5
0
0
50
100
n=4
150
x
200
250
300
FIG. 2: Protein number PDFs of our model, pn(x) =
pn(x; ω(n), b), vs. gamma distributions, pγ(x; α, β). Both
distributions have similar shapes and become identical for
ω = 1 (n = 1). Red solid line: pn(x) (20) for ω = 2n − 1,
n = 1, 2, 3, 4 and b = 5. Black dashed line: pγ(x; α, β) (12)
with α = 3ω/(ω + 2) and β/b = 2(ω + 2)/3 given by (13) for
M2 = 1/4. Blue lines: Simulations using the Gillespie algo-
rithm ([30], see Appendix K). Inset: Parameter comparison.
Because the protein copy number and not protein con-
centration is the variable in our model, the protein parti-
tioning at cell division along with an age structure of the
cell population (cf. Eq. (D2)) lead to the existence of the
noise floor -- an absolute lower bound for noise, present
for highly expressed proteins.
Our results suggest that the values of mean burst size
and burst frequency that may be obtained by fitting the-
oretical distributions to experimental data are strongly
model-dependent. Naïve fitting of the gamma distri-
bution [11] or its discrete counterpart, negative bino-
mial distribution [18], to the data measured in terms of
molecule numbers, would neglect protein loss at cell di-
vision because the underlying models neglected cell divi-
sion, and thus such a fitting might overestimate the mean
burst sizes and underestimate mean burst frequencies for
higher gene expression levels.
Our model, directly accounting for
the protein
molecule numbers, may be important for interpretation
of experimental results. In some studies, data correction
for cell size was carried out: In [12], cell volume was mea-
sured by image recognition and the protein fluorescence
was normalized by the volume. In other studies, the flow
cytometry data were gated [25, 26, 28, 29]. Gating filters
out the data only for those cells that scatter light to a
similar extent, so the observed cells are supposed to be of
similar sizes. However, the gating procedure may be im-
perfect because setting the size range too narrow leaves
too little data, whereas setting the gate too wide results
in some variation in cell sizes. One can avoid these prob-
lems by measuring the total cell fluorescence [3, 21, 27]
and then separating the noise contributions from varying
cell sizes (due to cell-cycle progression and other sources
of variability) using independent measurements [21] or
estimating these contributions theoretically. Our model
brings us closer to theoretical determination of the mag-
nitude of one of these contributions, namely, the protein
loss due to cell division and partitioning between daugh-
ter cells.
We have compared our model to the existing literature
data [21], however, the possibilities of such a comparison
are limited so far. The results for B. subtilis, reported in
Ref.
[21], were measured in too narrow a range of gene
expression levels, whereas the results for S. cerevisiae of
Refs.
[3, 27] may not be comparable to our minimal
model because of the complexity of gene expression mech-
anisms in eukaryotes. However, a cautious comparison of
our model to the above data suggests that the main lim-
itation of the model is its prediction of too large a con-
tribution of protein loss at cell division to the noise floor.
This is due to the exponential distribution of cell cycle
lengths, since the cell divisions are modelled as Poisso-
nian events (Appendix C). In a more realistic model, the
cell cycle length distribution could be modeled, instead
of the exponential distribution, as gamma distribution
[31] or some other distribution peaked around the mean
cell cycle length, with the limiting case of Dirac delta
distribution describing a deterministic cell cycle. How-
ever, this is beyond the scope of the present paper. In
preparation is our new paper [32] that explores the noise
levels in a more realistic model.
Another limitation of our theoretical approach may
be the description of the messenger RNA (mRNA) as
very short-lived molecules, only implicitly present in the
model [11], and thus the neglect of their partitioning
at cell division. Also, the model does not describe the
factors that may significantly shape the gene expression
noise levels in eukaryotes, e. g., the discrete on-off pro-
moter switching or nuclear transport.
We may expect that negative gene autoregulation (not
considered here) would suppress protein noise [11, 33],
and thus decrease the noise floor. Yet, positive autoregu-
lation is expected to have the opposite effect [11, 33]. In
a similar manner, nontrivial dependence of cell division
rate on protein number (hd(x) 6= const) would probably
result in a lower noise floor. Still, the analytical solution
to our model in presence of gene autoregulation or cell di-
vision regulation does not seem feasible. For that reason,
we have studied here the non-regulated gene expression
and cell division only.
Acknowledgments
AOM was supported by the National Science Centre
SONATA BIS 6 grant no. 2016/22/E/ST2/00558.
Author contributions
JJ designed the study, performed the analytical calcu-
lations and wrote the manuscript. MK wrote the simu-
7
lation code for cell division. AOM participated in the
study design by the idea of comparison of the model
with experimental data and by making the comparison;
designed the simulation algorithm for cell division, su-
pervised writing the simulation code and introduced mi-
nor modifications in the code; performed the simulations;
performed the distribution fitting by maxima matching
and observed heuristically the behavior of the probablity
distributions described in Appendix J; participated in
writing the manuscript.
Appendix A: Derivation of Eq. (1)
Equation (1) follows from an apparently very similar,
but more fundamental equation describing the number
density of cells, F (x, t), in a proliferating population,
∂F (x, t)
∂t
= −
∂
∂x
[g(x)F (x, t)] − hd(x)F (x, t)
+ 2Z 1
+ kZ x
0
0
η(q)
q
hd(cid:18) x
q(cid:19) F (cid:18) x
q
, t(cid:19) dq
w(x − x′)hp (x′) F (x′, t)dx′. (A1)
Namely, F (x, t)dx is the number of those cells in a
population, which at time t contain exactly x protein
molecules. In order to derive Eq. (1) from Eq. (A1), we
define
p(x, t) =
F (x, t)
N (t)
,
(A2)
where
N (t) = Z ∞
0
F (x, t)dx
(A3)
is the total number of cells in the population [15]. Inte-
grating both sides of Eq. (A1) from x = 0 to x = ∞,
assuming that F (∞, t) = g(0)F (0, t) = 0 and making use
of (A2) and (A3) we obtain
dN (t)
dt
= N (t)Z ∞
0
hd(x)p(x, t)dx.
(A4)
Now, Eq. (1) follows from (A1)-(A4) and from obvious
identity
∂F (x, t)
∂t
= N (t)
∂p(x, t)
∂t
+ p(x, t)
dN (t)
dt
.
(A5)
The above derivation is essentially that of Ref. [15], the
only difference is the presence of the terms responsible
for bursty protein production in the present case.
In spite of their apparent similarity, there are impor-
tant differences between Eq. (1) and Eq. (A1). First, the
terms related to protein partitioning in Eq. (A1) con-
serve the total number of protein molecules (molecules
are neither created, nor destroyed during cell division),
whereas from Eq.
(1) it follows that the cell division
always decreases the mean protein number in the pop-
0 hd(ξ)p(ξ, t)dξ
ulation. This is due to the (−1)p(x, t)R ∞
term in Eq. (1), not present in Eq. (A1).
Second, N (t) grows indefinitely according to Eq. (A4),
therefore in contrast to Eq. (1), Eq. (A1) does not have
nontrivial stationary solutions.
Our task now is to determine, for the present model:
(i) the population growth rate, νm, (ii) probability distri-
bution f (τ ) of generation time (cell cycle length), τ , for
each newborn cell, and (iii) the cell age (a) distribution
φ(a) in the state of a balanced, exponential growth [35].
This is done in next three sections.
Appendix B: Population growth rate νm
The steady state solutions of Eq.
(1), p(x), corre-
spond to the so called state of balanced growth, when
the shape of F (x, t) = N (t)p(x) does not change but is
only rescaled by N (t) [15]. In such a case, from Eq. (A4)
we obtain
N (t) = N0eνmt,
(B1)
where νm, given by
νm = Z ∞
0
hd(x)p(x)dx
(B2)
is the population growth rate. N (t) as given by (B1) is
characteristic for (in fact, defines) the phase of exponen-
tial growth [31][36].
If g(x) = −γx and if neither the protein production
rate nor the cell division rate depend on the number of
protein molecules (hp(x) = 1, hd(x) ≡ ∆/2), Eq.
(1)
reads
8
It is instructive to show this result in an alternative
way. Consider time evolution of the moments of p(x, t),
resulting from Eq. (B3),
µr = − [(1 − Mr)∆ + rγ] µr + k
r
Xl=1
(cid:18)r
l(cid:19)µr−lml. (B6)
In the absence of protein production (k = 0) and degra-
dation (γ = 0), from Eq. (B6) it follows that the time
dependence of the first moment of p(x, t), i.e., the mean
protein number is given by [37]
µ1(t) = µ1(0)e−∆0t.
(B7)
X(t)/N (t), where X(t) ≡ PN (t)
On the other hand, from the definition of the popula-
tion average, the mean protein number µ1(t) is equal to
i=1 xi(t) denotes the to-
tal number of protein molecules in a population, and
xi(t) is a number of protein molecules in i-th cell at
time t.
If there is no protein production or degrada-
tion, X(t) is constant and equal to its initial value,
X(t) = X(0) = N0µ1(0).
In such a case, time evolu-
tion of the mean protein number µ1(t) is caused solely
by the increase in cell number,
X(t)
N (t)
=
N0µ1(0)
N0eνmt = µ1(0)e−νmt.
(B8)
Comparing (B7) and (B8) we see that νm = ∆0, i.e., the
population growth rate νm is equal to the individual cell
division rate ∆0, as it should.
So far, we have considered the whole proliferating cell
population, for which ∆ = 2∆0 = 2νm (scenario A).
However, Eq. (B3), but not Eq. (1), may be also used
to describe the time evolution of the protein number dis-
tribution p(x, t) in a single cell lineage. In such a case,
we discard one of the daughter cells at each cell division,
and therefore ∆ = ∆0 (scenario B).
∂p(x, t)
∂t
= γ
∂
∂x
[xp(x, t)] + kZ x
η(q)(cid:20) 1
p(cid:18) x
q
q
0
+ ∆Z 1
0
w(x − x′)p(x′, t)dx′
, t(cid:19) − p(x, t)(cid:21) dq.
(B3)
Appendix C: Generation time distribution f (τ )
Central to this paper is the steady-state limit of Eq. (B3),
−γ
d
dx
[xp(x)] = ∆Z 1
+ kZ x
0
0
q(cid:19) − p(x)(cid:21) dq
p(cid:18) x
η(q)(cid:20) 1
w(x − x′)p(x′)dx′,
q
(B4)
(2). Although time-independent, Eq.
i.e., Eq.
(B4)
describes the stationary protein distribution p(x) in a
growing population of dividing cells (i.e., the state of
balanced growth), as discussed above. In particular, for
hd(x) = ∆/2 ≡ ∆0, from Eq. (B2) it follows that
νm = ∆0,
(B5)
i.e., the population growth rate νm appearing in Eq. (B1)
is equal to individual cell division rate ∆0, as should be
expected.
In order to find the generation time distribution f (τ ),
consider Eq. (B3). This is a special case of the differ-
ential Chapman-Kolmogorov equation [22] with x- and
t-independent coefficients. We assume once again that
there is no protein production (k = 0) and that the pro-
tein is stable (γ = 0). In such a case, (B3) becomes the
Master equation, describing 'jump process' between dif-
ferent states of the system (there is no drift term), and
these 'jumps' are solely due to the cell division events.
The probability that the system does not undergo such
a 'jump', and that it is still in the same state at t = τ
as it was at t = 0, is equal to exp(−∆τ ) [22]. Therefore,
the probability of a jump occurring in the infinitesimal
interval (t, t + dt) is given by [34]
π(t)dt = ∆ exp(−∆t)dt.
(C1)
In the case of a single lineage (scenario B mentioned
above), we have ∆ = ∆0, and π(t) must be identified with
the cell cycle length distribution, π(τ ) = f (τ ). Therefore,
we obtain
f (τ ) = ∆0 exp(−∆0τ ).
(C2)
If we deal with the whole proliferating population, then
∆ = 2∆0 (scenario A), and π(t) (C1) reads
π(t) = 2∆0 exp(−2∆0t).
(C3)
π(τ ) (C3) is identical with the quantity defined as
C(τ ) ≡ 2e−νmτ f (τ ) = 2∆0 exp(−2∆0τ ),
(C4)
and called the 'carrier distribution' in [31].
Appendix D: Cell age distribution φ(a)
To convince ourselves that (C2) is valid, and to find
the explicit form of the age distribution φ(a), assume
that f (τ ) = λ exp(−λt), for λ yet unspecified. From Eq.
(9) of Ref. [31] it follows that φ(a) is given by
φ(a) = 2νme−νm Z ∞
a
f (τ )dτ = 2νme−(νm+λ)a.
(D1)
R ∞
condition
The normalization of φ(a)
0 φ(a)da = 1) yields νm = λ, whereas from Eq.
(B5) or from (B7) and (B8) we obtain λ = ∆0, i.e., we
recover Eq.
(C2). Hence, the age distribution φ(a) is
given by
(i.e.,
the
φ(a) = 2∆0e−2∆0a.
(D2)
It should be emphasized that functional forms of f (τ )
(C2), C(τ ) (C4) and φ(a) (D2) are rather unrealistic.
This is a consequence of a Poissonian nature of cell di-
vision in the present model.
In particular, both f (τ )
and C(τ ) should be unimodal, and vanishing not only for
τ = 0, but also for the values of τ sufficiently close to zero
-- there certainly must be a minimal length of generation
time.
Because the functional forms of f (τ ) (C2) and C(τ )
(C4) are identical, one can treat the results of numeri-
cal simulation of a single cell lineage as referring to the
whole proliferating population, if only the division rate
is rescaled. Namely, for a single call lineage (scenario
B) we obtain identical protein number probability dis-
tribution as for the whole growing population (scenario
A), provided that in the latter case the the true division
rate ∆0 = νm ≡ ∆0(A) is two times smaller then in the
former, i.e., 2∆0(A) = ∆0(B).
Note that the simulation curves shown in Fig. 2 in the
main text were generated by simulation of a single cell lin-
eage with the cell division rate ∆0(B) (scenario B, see Ap-
pendix K). However, the theoretical curves pn(x; ω(n), b)
shown in Fig. 2 in the main text can be interpreted in
9
two ways, depending on how we define the ω parameter:
ω = k/∆0(A) assumes Scenario A (whole population) and
ω = k/∆0(B) = k/(2∆0(A)) assumes Scenario B (single
lineage). Therefore, the simulation results shown in Fig.
2 can also be reinterpreted as the results for Scenario A
(whole population) where the cell division rate was twice
smaller than the value set in the simulation algorithm:
∆0(A) = ∆0(B)/2.
The above discussion applies to the batch culture.
Analogous formulas can be derived for the continuous
cell culture [31], for which, in fact, conditions for state of
balanced growth can be more easily reached and main-
tained.
Appendix E: Noise floor: Overview of other
experimental results in the literature
A noise floor was observed in a number of experiments
reported in the literature, however, these data were not
suitable for comparison to our model. Below we overview
the existing studies, which we are aware of, and which
report the noise floor in gene expression.
In Ref.
[12], the noise floor in gene expression in E.
coli manifested itself as a boomerang-shaped log-log plot
of the squared coefficient of variation vs. mean gene ex-
pression. However, protein levels were measured in that
Reference as concentrations and not absolute molecule
numbers, which excluded the effect of protein loss at cell
division inherent to our model. The description of the
results in Ref. [12] may seem slightly misleading at first
glance because the plots in that reference were shown in
the units of protein numbers.
In fact, the protein flu-
orescence was measured in each cell and then its level
for that cell was normalized by the volume of that same
cell to get the protein concentration. The protein con-
centrations were then again normalized by a mean cell
volume to obtain the description in the units of molecule
numbers. And therefore, the resulting plots in Ref. [12]
show the protein numbers corresponding to the content
of average-sized cells, but the underlying method of mea-
surement intentionally removed the effects of protein loss
at cell division from the data. Thus, the results of Ref.
[12] should be interpreted using a model that describes
protein levels in terms of concentrations. For that reason,
the Friedman's model [11] describing protein concentra-
tions was used for data fitting in Ref.
[12]. Our model
cannot be fitted to these data because it describes protein
levels in terms of protein numbers and thus it explicitly
accounts for protein loss at cell division.
In Refs.
[28] and [26], a noise floor was observed in
gene expression in E. coli. These data were unsuitable
for comparison to our model because they were gated
to observe the gene expression levels in only those cells
that were in similar cell-cycle stages. Thus, the effect of
protein loss due to cell division would not be visible in
these data. A noise floor was also observed in gated data
in S. cerevisiae [25, 29].
10
For hd(x) = ∆/2 and hp(x) = 1, from (F7), (F8) and
(F9) we obtain the tail of the form exp{−x/[b(ω + 1)]}.
Note that the the above derivation is not universally
valid. For example, for the 'all or none' mode of pro-
tein partitioning, i.e., for ηb(q) = [δ(q) + δ(1 − q)] /2, one
cannot simply drop out the integral over q in Eq. (1).
However, apart from such rather pathological situations,
it seems that the derivation of the large-x asymptotic be-
haviour of p(x) proposed in Ref. [16] can be generalized
to the present case of stochastic protein production, pro-
vided the burst size PDF is of the form (F3) and there
are is no protein decay.
Appendix G: Distribution of protein numbers for
the uniform protein partition ratio distribution.
In this section we solve Eq. (B4) (Eq. (2) of the main
text) for γ = 0, exponential burst size PDF as given by
Eq. (F3), and for uniform partition ratio distribution
η(q) = Θ(q)Θ (1 − q) = 1,
(G1)
for 0 < q < 1, where Θ(q) denotes the Heaviside theta
function. η(q) (G1) is not very realistic at large x, where
any such distribution should be peaked around q = 1/2,
like, e.g., the symmetric beta distribution (a continuous
counterpart of the binomial distribution), considered in
Ref.
[15]. η(q) (G1) becomes more acceptable for x of
order of few protein molecules. Yet, η(q) (G1) has been
used in Refs.
[16, 20], probably due to its mathemati-
cal simplicity -- it is one of the few examples of a parti-
tion ratio PDF, for which analytical solutions of PBE-like
equations are known.
For γ = 0 and η(q) (G1), Eq. (3) reads
Appendix F: Distribution tails
The protein number distribution p(x) as given by Eqs.
(18) and (19) in the main text, i.e.,
p(x) =
∞
Xi=0
Ci(ω)
b
exp(cid:18) −2ix
b(ω + 1)(cid:19) ,
(F1)
where
Ci(ω) =
2iω
(2(1 + ω)−1; 2)i
( 1
2 (1 + ω)−1; 1
2 )∞
(1 + ω)2
(2; 2)i
( 1
2 ; 1
2 )∞
,
(F2)
has exponential tail of the form exp{−x/[b(ω + 1)]} (the
leading exponent in (F1)). The same is true for pu(x)
(G6) (see the next section), i.e. the solution of Eq. (B4)
for η(q) = Θ(q)Θ (1 − q) = 1 given by Eq.
In
fact, it can be shown that for most of functional forms
of η(q), the solution of Eq. (B4) has exactly the same
exponential tail (there are, however, exceptions: for ex-
ample, η(q) = 1
2 [δ(q) + δ(1 − q)] yields the tail of the
form exp{−x/[b(2ω + 1)]}).
(G1).
This is a special case of a yet more general result. We
assume here that g(x) = 0 and the burst size PDF is
exponential, as given by Eq. (10),
ν(u) =
e−u/b
b
,
L[ν(u)] ≡ ν(s) = 1/(sb + 1),
(F3)
(1). Following Ref.
but we allow for arbitrary x-dependence of hd(x) and
hp(x) in Eq.
[16], we neglect the
q-integral term in the steady-state limit of Eq. (1) (for
large protein number x, the contribution of states with
still larger x may be neglected) and obtain
0 = −(cid:20)hd(x) +Z ∞
+ kZ x
0
0
w(x − x′)p(x′)dx′.
hd(ζ)p(ζ)dζ(cid:21) p(x)
Next, we combine the derivative of (F4) with respect to
x with the original equation and obtain
kh′
p(x) + h′
d(x) + hd(x)/b + R/b
khp(x) + hd(x) + R
p′(x)
p(x)
=
where
R = Z ∞
0
hd(ζ)p(ζ)dζ = νm,
(F7)
c.f. Eq. (B2). From (F6) it follows that at the tail of the
PDF the ratio p(x2)/p(x1) is well approximated by
P12 = exp (−I12) ,
(F8)
(F4)
0 = k w(s)G(s) + ∆Z 1
0
[G(qs) − G(s)]dq.
(G2)
By differentiating (G2) with respect to s and combining
such obtained equation with the original one, integrating
by parts and using the identity
,
(F5)
(F6)
Z 1
0
qG′(qs)dq =
1
s (cid:20)G(s) −Z 1
0
G(qs)dq(cid:21) ,
(G3)
in order to to cancel the terms containing G(qs) (note
that G′(qs) ≡ (dG(y)/dy)y=qs), we finally get
G′(s) =
ω [s w′(s) + w(s)]
s[1 − ω w(s)]
G(s).
(G4)
ODE (G4) is equivalent to the integral equation (G2).
For ν(s) (F3) we have w(s) = ν(s) − 1 = −sb/(sb + 1)
and Eq. (G4) has the following solution:
where
I12 = Z x2
x1
kh′
p(x) + h′
d(x) + hd(x)/b + R/b
khp(x) + hd(x) + R
dx.
(F9)
G(s) = (1 − ǫ)
ξǫ
(s + ξ)ǫ + ǫ
ξǫ+1
(s + ξ)ǫ+1 ,
(G5)
where ǫ = ω/(ω + 1), ξ = 1/[b(ω + 1)]. The inverse
Laplace transform of G(s) (G5) reads
pu(x) = (1 − ǫ)
ξǫxǫ−1e−ξx
Γ(ǫ)
+ ǫ
ξǫ+1xǫe−ξx
Γ(ǫ + 1)
.
(G6)
Because pu(x) (G6) is a statistical mixture of two broad
gamma distributions (note that 0 < ǫ < 1), pu(x) itself
is broad (c2
v ≥ 1/2).
For the case of deterministic protein production, in-
stead of pu(x) (G6) we obtain a gamma distribution
[16, 20], cf. next Section.
Appendix H: Distribution of protein numbers for
the deterministic protein production and
half-by-half protein partitioning
We consider here the case of unregulated, deterministic
protein production, i.e., in Eq. (1) we put k = 0 and
g(x) = σ > 0. For hd(x) = ∆/2, in the steady-state limit
instead of Eq. (B4) (Eq. (2)) we obtain
λ
d
dx
[p(x)] = Z 1
0
η(q)(cid:20) 1
q
p(cid:18) x
q(cid:19) − p(x)(cid:21) dq,
(H1)
where λ = σ/∆. Laplace transform of Eq. (H1) yields
λsG(s) = Z 1
0
η(q) [G(qs) − G(s)] dq,
(H2)
and we have assumed p(0+) ≡ limx→0+ p(x) = 0 (validity
of this assumption is checked a posteriori). For the half-
by-half protein partitioning, η(q) = δ(q − 1
2 ), Eq. (H2)
can be rewritten as
G(s) = (1 + λs)
−1 G(cid:16) s
2(cid:17) .
(H3)
Eq. (H3) can be solved by iteration; we obtain
G(s) =
∞
Yi=0
1
1 + 2−iλs
=
1
(−λs; 1
2 )∞
=
∞
Xr=0
where (a; q)k is a q-Pochhammer Symbol.
have used the following identity
1
(z; q)∞
=
∞
Xn=0
zn
(q; q)n
,
(−λs)r
( 1
2 ; 1
2 )r
,
(H4)
In (H4) we
(H5)
which is a special case of the q-binomial theorem [24].
Note that G(s) (H4) can be also rewritten as
G(s) =
∞
Xr=0
(−2λs)r
[r] 1
2
!
= e 1
2
(−2λs),
(H6)
where [r]q! ≡ (1 − q)−r(q; q)r denotes the q-factorial and
eq(x) is the q-exponential function.
11
0.5
0.5
0.4
0.4
)
)
λ
λ
;
;
x
x
(
(
p
p
0.3
0.3
,
,
)
)
λ
λ
;
;
x
x
(
(
γ
γ
p
p
0.2
0.2
0.1
0.1
0
0
0
0
1
1
2
2
3
3
4
4
x
x
5
5
6
6
7
7
8
8
FIG. 3: Protein number PDFs of our model in case of deter-
ministic protein production. p(x; λ) (H7) (red) is a solution
of Eq. (H2) for η(q) = δ(q − 1/2) (9), whereas the gamma
distribution pγ(x; λ) = λ−2x exp(−x/λ) (H11) (green) is a so-
lution of Eq. (H2) for η(q) = Θ(q)Θ (1 − q) (G1); both are
plotted here for λ = 1, i.e., as a function of the rescaled vari-
able x/λ. Both distributions have the same average value,
hxi = µ1 = 2λ, and identical exponential tails ∼ exp(−x/λ)
but seemingly different shapes. The variance of p(x; λ), equal
3 λ2, is smaller than the variance of pγ(x; λ), equal to 2λ2.
to 4
In contrast to pγ (x; λ), the maximum of p(x; λ) cannot be
for λ = 1 we have xm ≈ 1.2773 and
found analytically;
p(xm) ≈ 0.4549. Note that, for both p(x; λ) and pγ(x; λ), the
squared coefficient of variation is constant (does not depend
on λ, i.e., on the mean protein number), and is equal to its
minimal value for the corresponding solution with stochastic
protein production.
The inverse Laplace transform of (H4) reads
p(x; λ) =
1
λ
∞
Xi=0
Di exp(cid:18)−
2ix
λ (cid:19) ,
(H7)
where the coefficients Di are given by
Di =
i
Yj=1
2
1 − 2j
∞
Yl=1
2l
2l − 1
=
2i
(2; 2)i
1
2 ; 1
( 1
2 )∞
.
(H8)
For i = 0, we obtain the leading-order term, i.e., the
slowest-decaying exponent, e−x/λ, which determines the
behaviour of p(x; λ) (H7) at the x → ∞ limit. Note
that p(x; λ) (H7) can be regarded as a counterpart of
the probability distribution given by (F1) and (F2) (Eqs.
(14) and (15) in the main text) in the case of determin-
istic protein production.
From G(s) (H4) we obtain both the moments (µr) and
the cumulants (κr) of L−1[G(s)] ≡ p(x; λ) (H7), namely
µr =
λrr!
2 ; 1
( 1
2 )r
=
2rλrr!
!
[r] 1
2
,
κr =
2rλr(r − 1)!
2r − 1
.
(H9)
(H10)
In the present case, µr can be also easily obtained from
moments equations.
For a uniform distribution of protein division ratio
(G1), the solution of Eq. (H1) reads [16]
pγ(x; λ) = λ−2xe−x/λ.
(H11)
The gamma distribution pγ(x; λ) (H11) is therefore a de-
terministic counterpart of pu(x) (G6).
Both p(x; λ) (H7) and pγ(x; λ) (H11) define a one-
parameter family of probability distributions; we have
p(x, λ) = λ−1f(x/λ), where f (z) = p(z, 1) is a func-
tion of the rescaled variable x/λ. Also, both p(x; λ)
and pγ(x; λ) are right-skewed, unimodal, and vanish in
the x → 0+ limit, as has been assumed for p(x; λ), cf.
Fig 3. However, in contrast to the gamma distribution,
for p(x; λ) we have p(k)(0; λ) = 0 for arbitrary k, hence
p(x; λ) is not analytic at x = 0. In consequence, the Tay-
lor expansion of p(x; λ) at x0 = 0 does not exists. This
can be shown using properties of G(s) (H4). We have
p(0) = lim
s→∞
p′(0) = lim
s→∞
p′′(0) = lim
s→∞
. . .
sG(s) = 0,
s [sG(s) − p(0)] = 0,
s(cid:2)s2G(s) − sp(0) − p′(0)(cid:3) = 0,
(H12)
etc., due to the obvious relation
lim
s→∞
skG(s) = 0,
(H13)
valid for any k < ∞.
12
n=2
n=3
n=4
;
(n)
;
FIG. 4: Gamma distributions pγ (x; α, β) (I2) are almost indis-
tinguishable from the pn(x; ω(n), b) distributions of our model
as given by Eq. (I1), when fitted by matching their maxima.
Red lines: pn(x; ω(n), b) with b = 5 and n = 2, 3, 4. Green
lines: The corresponding pγ(x; α, β) (I2), whose parameter
values are given in Table I.
The coordinates xmax,pn and ymax,pn of the maximum of
pn(x; ω(n), b) can be found analytically for small values
of n (in our case, the Maple software was able to find
them explicitly for n ≤ 5). Otherwise, they can be calcu-
lated numerically. In Fig. 4, we put b = 5 and n = 2, 3, 4
(corresponding to ω = 3, 7, 15) in pn(x; ω(n), b). By nu-
merically solving the equations
Appendix I: Similarity of gamma distributions and
the distributions resulting from our model
xmax,pn = xmax,pγ ,
ymax,pn = ymax,pγ ,
(I4)
In this section we present an alternative method of fit-
ting of gamma distributions to the distributions resulting
from our model: We match the maxima of our distribu-
tion,
n
1
b
x
(I1)
pn(x; ω(n), b) =
2lb(cid:17) ,
An,l exp(cid:16)−
i=1(2i − 1)Qn−l
where An,l = (−1)n−l2
j=1(2j − 1)]
(c.f. Eqs.
(20) and (21) of the main text) and of the
gamma distribution pγ(x; α, β), c.f. Eq. (12) of the main
text, i.e.:
Xl=1
2 /[Ql−1
l(l−3)
pγ(x; α, β) ≡
xα−1 exp (−x/β)
βαΓ(α)
.
(I2)
The coordinates of the gamma distribution's maximum
for α > 1 are given by:
xmax,pγ = β α − β,
ymax,pγ =
[(α − 1) β]α−1 exp(1 − α)
βαΓ (α)
.
(I3)
we found the α and β parameters of the gamma dis-
tributions whose maxima exactly match the maxima of
pn(x; ω(n), b).
Fig.
4 shows that this way of fitting yields the
gamma distribution plots that are even more similar to
pn(x; ω(n), b) than those obtained by matching the first
two moments of the distributions, as in the main text,
Fig 2.
n
ω
α
β/b
2
3
1.90
3.07
3
7
2.52
5.28
4
15
2.88
9.74
TABLE I: Parameters of the distributions pn(x; ω(n), b) and
pγ(x; α, β) shown in Fig. 4.
Appendix J: A property of the distribution pn(x)
We denote the Laplace transform of pn(x) (I1) as
Gn(s). From Eq. (12) of the main text, it follows that
Gn(s) =
One can show that
(cid:0)s + b
4(cid:1) . . .(cid:0)s + b
2n(cid:1)
bn
2
2− n(n+1)
2(cid:1)(cid:0)s + b
s Gn+1(s) =
b
2n+1 [Gn(s) − Gn+1(s)] .
Assuming
pn+1(0) = 0 for n > 0,
it follows from Eq. (J2) that
.
(J1)
(J2)
(J3)
(J4)
d
dx
pn+1(x) =
b
2n+1 [pn(x) − pn+1(x)] .
(J4) iteratively for n ≥ 1 to ob-
We can solve Eq.
tain the formulas for consecutive pn+1(x), starting from
p1(x) = (b/2) exp(−bx/2) and using (J3) as the bound-
ary condition. From Eq. (J4) it follows that pn+1(x) has
a maximum in the point where its plot intersects with
the plot of pn(x).
Appendix K: Simulation
The simulation results shown in Fig. 2 in the main text
were obtained using a custom extension of the StochPy
package [30]: The standard simulation using the direct
Gillespie algorithm was supplemented with cell division.
Histograms were calculated along a single trajectory,
which mimics the observation of a single cell lineage. The
initial part of the trajectory was cut off to obtain only the
steady-state behavior. The reaction kinetics (see the file
NonRegulatedGeneCellDivision.psc) was given by:
13
where O denotes the gene promoter, M is mRNA, and
P is protein. k is transcription rate and kp is transla-
tion rate. kdm denotes mRNA degradation rate and its
value is chosen such that mRNA life time is much shorter
than the mean cell cycle duration 1/∆0(B), consistently
with the assumption of the analytical model described in
the main text. Additionally, cell division occurs at the
rate ∆0(B): The custom function multiple_division()
draws a random cell cycle length T from a specified dis-
tribution (here: exponential with the mean 1/∆0(B)).
In a given cell cycle, the simulation runs until the cell
age T and at that time point the molecule numbers M
and P are divided by 2.
If the remainder of the divi-
sion is 1, then the remaining molecule goes to the "ob-
served" daughter cell with probability 1/2. The simula-
tion for the next cell cycle is initialized with the result-
ing molecule numbers M and P for the daughter cell.
Simulation parameter values are shown in Table II and
III. The simulation code, containing the custom func-
tions that implement cell division can be found in the file
NonRegulatedGeneCellDivision.ipynb (Jupyter note-
book file, Supplemental Material) [38].
∆0(B)
kp
kdm
max_simulation_time
time_cutoff
1
250
50
20000
2000
TABLE II: Simulation parameter values for Fig. 2 in the main
text. These values were the same in all simulations.
n 1
1
k
2
3
3
7
4
15
O k−−→ O + M
kp−−→ M + P
M
kdm−−→ ∅
M
(K1)
TABLE III: Simulation parameter values for Fig. 2 in the
main text. These values correspond to the specific curves in
Fig. 2.
[1] M. Hintsche and S. Klumpp, Journal of Biological Engi-
11 (2016).
neering 7, 22 (2013).
[2] S. Klumpp and T. Hwa, Current Opinion in Biotechnol-
ogy 28, 96 (2014).
[6] D. Huh and J. Paulsson, Nature Genetics 43, 95 (2011).
[7] D. Huh and J. Paulsson, Proceedings of the National
Academy of Sciences 108, 15004 (2011).
[3] L. Keren, D. Van Dijk, S. Weingarten-Gabbay, D. Davidi,
G. Jona, A. Weinberger, R. Milo, and E. Segal, Genome
Research 25, 1893 (2015).
[8] R. Marathe, V. Bierbaum, D. Gomez, and S. Klumpp,
Journal of Statistical Physics 148, 608 (2012).
[9] J. Paulsson, Physics of Life Reviews 2, 157 (2005).
[4] V. Shahrezaei and S. Marguerat, Current Opinion in Mi-
[10] L. Cai, N. Friedman, and X. S. Xie, Nature 440, 358
crobiology 25, 127 (2015).
(2006).
[5] N. Walker, P. Nghe, and S. J. Tans, BMC Biology 14,
[11] N. Friedman, L. Cai, and X. S. Xie, Physical Review
14
Letters 97, 168302 (2006).
441, 840 (2006).
[12] Y. Taniguchi, P. J. Choi, G.-W. Li, H. Chen, M. Babu,
and X. S. Xie, Science 329, 533
J. Hearn, A. Emili,
(2010).
[13] M. R. Maurizi, Experientia 48, 178 (1992).
[14] N. V. Mantzaris, Computers & Chemical Engineering 29,
897 (2005).
[15] N. V. Mantzaris, Journal of Theoretical Biology 241, 690
(2006).
[16] T. Friedlander and N. Brenner, Physical Review Letters
101, 018104 (2008).
[26] L. Wolf, O. K. Silander, and E. van Nimwegen, eLife 4,
1 (2015).
[27] D. Volfson, J. Marciniak, W. J. Blake, N. Ostroff, L. S.
Tsimring, and J. Hasty, Nature 439, 861 (2006).
[28] O. K. Silander, N. Nikolic, A. Zaslaver, A. Bren,
I. Kikoin, U. Alon, and M. Ackermann, PLoS Genet-
ics 8, e1002443 (2012).
[29] A. Bar-Even, J. Paulsson, N. Maheshri, M. Carmi,
E. O'Shea, Y. Pilpel, and N. Barkai, Nature Genetics
38, 636 (2006).
[17] J. Yu, J. Xiao, X. Ren, K. Lao, and X. S. Xie, Science
[30] T. R. Maarleveld, B. G. Olivier, and F. J. Bruggeman,
311, 1600 (2006).
PloS one 8, e79345 (2013).
[18] V. Shahrezaei and P. S. Swain, Proceedings of the Na-
tional Academy of Sciences 105, 17256 (2008).
[19] J. Jedrak and A. Ochab-Marcinek, Physical Review E
[31] E. Powell, Microbiology 15, 492 (1956).
[32] J. Jędrak and A. Ochab-Marcinek, (2018).
[33] C. Jia, P. Xie, M. Chen, and M. Q. Zhang, Scientific
94, 032401 (2016).
Reports 7, 16037 (2017).
[20] N. Brenner and Y. Shokef, Physical Review Letters 99,
[34] N. van Kampen, Stochastic Processes in Physics and
138102 (2007).
[21] N. Nordholt, J. van Heerden, R. Kort, and F. J. Brugge-
man, Scientific Reports 7, 6299 (2017).
[22] C. Gardiner, Stochastic methods (Springer Berlin, 2009).
[23] F. Jafarpour, C. S. Wright, H. Gudjonson, J. Riebling,
E. Dawson, K. Lo, A. Fiebig, S. Crosson, A. R. Dinner,
and S. Iyer-Biswas, Physical Review X 8, 021007 (2018).
[24] G. E. Andrews, q-Series: Their development and applica-
tion in analysis, number theory, combinatorics, physics,
and computer algebra, Vol. 66 (American Mathematical
Society, 1986).
[25] J. R. Newman, S. Ghaemmaghami, J. Ihmels, D. K. Bres-
low, M. Noble, J. L. DeRisi, and J. S. Weissman, Nature
Chemistry (Elsevier Science, Amsterdam, 2007).
[35] In this section we use notation of Ref. [31], but the cell age
is denoted a and not a to avoid confusion with quantity
a = k/γ defined in the main text.
[36] We assume that the number of cells in the population,
N (t),
is sufficiently large its fluctuations may be ne-
glected, and that N (t) may be regarded as evolving ac-
cording to deterministic equation (B1).
[37] In such a case the only possible steady state is a trivial
one, i.e., limt→∞ p(x, t) = p(x) = δ(x).
[38] See Supplemental Material at [URL will be inserted by
publisher] for the simulation code.
|
1704.07407 | 1 | 1704 | 2017-04-24T18:38:12 | Exploring A Multi-Scale Method for Molecular Simulations in Continuum Solvent Model: Explicit Simulation of Continuum Solvent As An Incompressible Fluid | [
"physics.bio-ph",
"physics.chem-ph",
"physics.comp-ph"
] | A multi-scale framework was recently proposed for more realistic molecular dynamics simulations in continuum solvent models by coupling a molecular mechanics treatment of solute with a fluid mechanics treatment of solvent, where we formulated the physical model and developed a numerical fluid dynamics integrator. In this study, we incorporated the fluid dynamics integrator with the Amber simulation engine to conduct atomistic simulations of biomolecules. At this stage of the development, only nonelectrostatic interactions, i.e., van del Waals and hydrophobic interactions are included in the multi-scale model. Nevertheless numerical challenges exist in accurately interpolating the highly nonlinear van del Waals term when solving the finite-difference fluid dynamics equations. We were able to bypass the challenge rigorously by merging the van del Waals potential and pressure together when solving the fluid dynamics equations and by considering its contribution in the free-boundary condition analytically. The multi-scale simulation engine was first validated by reproducing the solute-solvent interface of a single atom with analytical solution. Next, we performed the relaxation simulation of a restrained symmetrical monomer and observed a symmetrical solvent interface at equilibrium with detailed surface features resembling those found on the solvent excluded surface. Four typical small molecular complexes were then tested, both volume and force balancing analysis showing that these simple complexes can reach equilibrium within the simulation time window. Finally, we studied the quality of the multi-scale solute-solvent interfaces for the four tested dimer complexes and found they agree well with the boundaries as sampled in the explicit water simulations. | physics.bio-ph | physics | Exploring A Multi-Scale Method for Molecular Simulations in
Continuum Solvent Model: Explicit Simulation of Continuum
Solvent As An Incompressible Fluid
Li Xiao1,2 and Ray Luo1,2,3
1. Departments of Biomedical Engineering, 2. Molecular Biology and Biochemistry, and 3.
Chemical Engineering and Materials Science, University of California, Irvine, CA 92697
A multi-scale framework was recently proposed for more realistic molecular dynamics
simulations in continuum solvent models by coupling a molecular mechanics treatment of solute
with a fluid mechanics treatment of solvent, where we formulated the physical model and
developed a numerical fluid dynamics integrator. In this study, we incorporated the fluid
dynamics integrator with the Amber simulation engine to conduct atomistic simulations of
biomolecules. At this stage of the development, only nonelectrostatic interactions, i.e., van del
Waals and hydrophobic interactions are included in the multi-scale model. Nevertheless
numerical challenges exist in accurately interpolating the highly nonlinear van del Waals term
when solving the finite-difference fluid dynamics equations. We were able to bypass the
challenge rigorously by merging the van del Waals potential and pressure together when solving
the fluid dynamics equations and by considering its contribution in the free-boundary condition
analytically. The multi-scale simulation engine was first validated by reproducing the solute-
solvent interface of a single atom with analytical solution. Next, we performed the relaxation
simulation of a restrained symmetrical monomer and observed a symmetrical solvent interface at
equilibrium with detailed surface features resembling those found on the solvent excluded
surface. Four typical small molecular complexes were then tested, both volume and force
balancing analysis showing that these simple complexes can reach equilibrium within the
simulation time window. Finally, we studied the quality of the multi-scale solute-solvent
interfaces for the four tested dimer complexes and found they agree well with the boundaries as
sampled in the explicit water simulations.
Please send correspondence to: [email protected]
1
1. Introduction
Atomistic simulation has become an important tool for studying the structures, dynamics, and
functions of biomolecular systems. Nevertheless efficient atomistic simulation of large and
complex biomolecular systems is still one of the remaining challenges in computational
molecular biology. The computational challenges in atomistic simulation of biomolecular
systems are direct consequences of their high dimensionalities. Indeed biomolecules are highly
complex molecular machines with thousands to millions of atoms. What further complicates the
picture is the need to realistically treat the interactions between biomolecules and their
surrounding water molecules that are ubiquitous and paramount important for their structures,
dynamics, and functions.
To appreciate these challenges, it is instructive to highlight the two bottlenecks in
biomolecular simulations: (1) the cost of each energy evaluation that is determined by the
number of particles in a mathematical model; and (2) the number of time steps of dynamics that
it takes for sufficient coverage of different conformations. Hundreds of millions of time steps are
routinely required in biomolecular simulations to draw statistically significant conclusions. It is
often the case that more particles need more time steps for sufficient coverage. Thus the overall
simulation cost usually scale exponentially with the number of particles in the mathematical
model used in a simulation. Indeed many fundamental and interesting biomolecular processes
remain largely inaccessible to atomistic simulations when system sizes exceed more than a few
hundred residues.
Since most particles in biomolecular simulations are to represent water molecules
solvating the target biomolecules, an implicit treatment of water molecules allows greatly
increased simulation efficiency. Indeed, implicit solvation offers a unique opportunity for more
2
efficient simulations without the loss of atomic-level resolution for biomolecules [1-17].
Advance in implicit solvation, coupled with developments in sampling algorithms, classical
force fields, and quantum approximations, will prove useful to the larger biomedical community
in a broad range of studies of biomolecular structures, dynamics and functions.
One class of implicit solvent models, the classical Poisson-Boltzmann (PB) solvent
model, has become widely accepted in biomolecular applications after over 30 years of basic
research and development. Efficient numerical PBE-based solvent models have been widely used
to study biological processes including predicting pKa values [18-21], computing solvation and
binding free energies [22-31], and protein folding [32-42]. However, challenges remain to
achieve more consistent, accurate, and robust analysis of biomolecules [43-60]. The existing
dielectric model based on molecular solvent excluded surface is a major hurdle for applications
of the Poisson-Boltzmann solvent models. This dielectric model is ad-hoc, expensive, and
numerically unstable due to its treatment of atoms as hard spheres in molecular simulations.
In our previous study [61], we explored a multi-scale simulation strategy to explicitly
simulate the continuum solvent/solute interface with the solvent fluid dynamics that is coupled to
the solute molecular dynamics. This strategy (1) allows a self-consistent treatment of the
solvation interactions, i.e. the dielectric interface automatically adjusts to local conformational
and energetic fluctuations and is guaranteed to be at the system free energy minimum upon
equilibrium; (2) allows a "soft" and more physical dielectric interface for stable dynamics; (3)
eliminates atom-specific cavity radii that must be defined, dramatically reducing the freely
adjustable parameters of the continuum solvent treatment; (4) eliminates the expensive molecular
surface reconstruction step during dynamics; and (5) eliminates the difficult and expensive
molecular surface-to-atom mapping of dielectric boundary forces and hydrophobic boundary
3
forces, and applies these surface forces to the continuum solvent instead. In addition, a 3D
numerical algorithm was developed to simulate the implicit solvent via the Navier-Stokes
equation [61, 62]. It should be pointed out that the use of Navier-Stokes equation, instead of
Stokes equation that is sufficient for biomolecular processes of interest, is necessary for the lack
of a predefined solute-solvent interface, or in a "free boundary" problem [61, 62]. Our numerical
algorithm was validated with multiple model test cases, demonstrating its effectiveness and
numerical stability, with observed accuracy consistent with the designed numerical algorithm.
In this study, we intended to explore the feasibility of incorporating the fluid dynamics
algorithm into the Amber molecular mechanics simulation engine [63-65] to assess the feasibility
and quality of the new multi-scale model for potential applications to biomolecular simulations.
At the current stage, we are mainly interested in equilibrium properties of the biomolecular
solute, and solvent hydrodynamics is not our consideration. Thus certain alterations of the
original model can be utilized to artificially accelerate the solvent relaxation process so that the
precious computing resources can be focused on sampling of solute conformations.
2. Theoretical Model
In the following we first review our physical model for easy understanding of the overall
approach. Next we briefly go over the fluid dynamics algorithm and procedure with a focus on
what has been revised from our previous study to adapt the method to atomistic molecular
simulations. Finally computational details are presented for the numerical tests of specific
molecular systems.
2.1 Physical model
Our basic model is derived from the Hamiltonian equation. A Hamiltonian for the entire system
is thus defined first. Its degrees of freedom are atomic positions (x ) and their velocities (v ) for
4
the solute molecular dynamics (MD) region; and fluid element displacements ( y ) and their
velocities (u) for the solvent fluid dynamics (FD) region. For the MD region, all-atom molecular
mechanics will be used. Molecular mechanics usually adopts a relatively simple potential energy
function, or force field, for efficient computation. Many potential energy functions have been
developed for biomolecular applications, such as Amber [66-71], CHARMM [72-74] and OPLS
(1)
[75-77]. For FD region, an incompressible viscous fluid model is adopted.
The Hamiltonian is defined as
H = H MD(x,αx) + H FD(y,αy) + H MD/FD(x,αx;y,αy),
where αx is the momentum of MD region and αy is the momentum of FD region. H MD is the
Hamiltonian of for the MD region modeled by molecular mechanics H MD =U + K , where U is
the force field potential energy and K is the kinetic energy. H FD represents the Hamiltonian for
the incompressible solvent fluid. HMD/FD =Uele +Uvdw +Uhse , consists of three terms. Uele is the
Poisson-Boltzmann electrostatic solvation energy [78-81]. The nonelectrostatic solvation energy
is modeled as two components: the van del Waals component Uvdw and the hard sphere
entropy/cavity component Uhse [82-86]. Here Uele is defined as
⎞
D⋅E − ΔΠλ
⎠⎟
1
8π
Uele = ρfϕ−
⎛
∫
⎝⎜
∑ (e−qiϕ/kT −1)
ΔΠ = kT
ci
i
dv
(2)
and Uvdw and Uhse are defined as
ρaw(raw)
Ns∑
∫a=1
Uvdw =
Uhse =γ i SAS + c
uLJ(raw)draw
(3)
5
Here the sum is over all solute atoms ( Ns ), and the integration is over the solvent-occupied
volume. ρaw(raw) is a solvent distribution function around solute 'a' at a given solute-solvent
distance. uLJ(r) =
A
r12 −
B
r6 is the force field Lennard Jones potential given the coefficient A, B
for each atom. γ is the surface tension and c is an offset constant.
Now we proceed to derive the dynamics equation by first setting β= (x,y) as the
position vector of the system and α= (αx,αy) as the momentum vector of the system. The
familiar Newtonian dynamics can be derived from the Hamilton's equation
.
.
α
= − ∂H
∂β
(4)
(5)
Here we have adopted the convention that α and β represent the moment and position vectors
of each particle/element, respectively.
In the molecule dynamics region, the equation of motion for an atom at position vector x
can be expressed symbolically as
.
αx
= − ∂H MD
∂x − ∂H MD/FD
∂x
.
− ∂H MD
∂x
represents the usual force field terms in molecule dynamics simulations. The coupling
Hamiltonian has three terms, Uele +Uvdw +Uhse . Since Uhse does not depend on atomic positions,
the coupling force terms that the atoms feel are only those of electrostatics and van del Waals in
nature, i.e., − ∂H MD/FD
∂x
∂x . It is interesting to note that the electrostatic forces
− ∂Uele
∂x are simply the qE forces, where q's are the "free" charges, i.e. atomic point charges in a
= − ∂Uele
∂x − ∂Uvdw
6
.
u.
(6)
(7)
force field model [87]. − ∂Uvdw
∂x
modeled as continuum [84] .
are the van del Waals forces from the solvent molecules
In the fluid dynamics region, consider a small fluid volume element at position y, with
volume V and velocity u. The equation of motion of the fluid element is
= − ∂H FD
∂y − ∂H MD/FD
∂y
i
αy
As shown below, the variational principle will be applied on this element. The partial derivative
can also be written in the variational form as
= −δH FD
δy
.
αy
u − δH MD/FD
δy
Here the subscript u denotes that it is fixed during the variation. Notice here y = y(b,t) is the
Lagrangian coordinate of the volume element, which is fixed on the fluid element and b is
introduced here to denote the actual spatial position [88].
Let us first focus on the variation of H FD , which has the form of
H FD =
1
2 ρu2 +Uint(ρ,s)
∫
⎡
⎣⎢
⎤
⎦⎥dV
, (8)
where Uint is the internal energy density, ρ represents the fluid density, and s is entropy. For a
small fluid volume element at position y and volume V , we impose a variation δy on the
element within time dt , and proceed to compute the variation of H FD . The process is assumed
δy
dt 0 . Since the fluid is incompressible, ∇ iδy = 0 , the work done to the
to be very rapid, i.e.,
environment is
dW = pδy i dS
∫
= ∇ i(pδy)
∫
∫
dV = ∇p iδy
dV + p∇ iδy
dV = ∇p iδy
dV, (9)
∫
∫
7
where p is pressure and vector dS denotes the surface element of the element with the direction
along the normal direction of the surface. Given that the first and second laws of
thermodynamics still hold, the internal energy variation can be expressed as
= (Tδs)
∫
The entropy constraint also gives
δUint dV
∫
dV − dW = (Tδs)
dV − ∇p iδy
dV. (10)
∫
∫
∫
(Tδs)dV
= fvis iδydV
∫
, (11)
where q is the heat flux and fvis is the viscous force density. Substitution of eqn (11) into eqn
(10) and the fact that the term involving dt can be ignored as δy
given the variation of the
− dt ∇ iqdV
∫
dt 0
internal energy as
δUint dV
∫
= (fvis iδy)
∫
∫
dV − ∇p iδy
dV = (f iδy)
dV, (12)
∫
where
the
total
force
density
f = fvis − ∇p =
∂σij
∂yj
is
introduced
and
⎛
σij = −pδij + µ
⎝⎜
∂(u⋅yi)
∂yj
+
∂(u⋅yj)
∂yi
⎞
⎠⎟ is the stress tensor and µ is the fluid viscosity constant [89].
of the fluid [89]. Given the assumption that the force density is uniform within the volume
element, substitution of eqn (8) and eqn (12) into the variation of H FD gives
−δH FD
δy
u = −
∫
δUint dV
δy
= −fV = −
∂σij
∂yj
V. (13)
The variation of HMD/FD is presented next. In Poisson-Boltzmann systems with mobile
ions, there is an ionic force term at the Stern layer [87], but it is usually much smaller than other
force terms, and is often ignored. If it were not ignored, the ionic force would act upon relevant
volume elements. Uhse only depends on the interface boundary so that it does not change under
8
the variation of the volume element. Thus the only significant derivative of HMD/FD is the van del
Waals force, which can be treated as the "external force" density (F ) on the fluid element, i.e.
−δH MD/FD
δy
u = −δUvdw
δy = FV. (14)
Finally, the change of momentum of the fluid volume element is
d(ρVu)
dt
=
i
αy
= ρV ∂u
∂t + ρV ∂u
∂ai
∑
i
∂ai
∂t
= ρV ∂u
∂t + ρV(u i ∇)u.
(15)
Combination of eqns (6) and (13) –(15) gives
ρV ∂u
⇒ ρ∂u
∂t + ρV(u i ∇)u =
∂t + ρ(u i ∇)u =
∂σij
∂yj
∂σij
∂yj
V + FV
+ F.
(16)
Including the conservation of volume/mass for the given volume element, i.e. ∇ iu = 0 , the
incompressible Navier-Stokes equation can be expressed as
ρ ∂u
⎛
⎝⎜
∂t +(u⋅∇)u
⎞
⎠⎟ = −∇p + µΔu + F
∇⋅u = 0
. (17)
2.2 Derivation of interface conditions
To obtain the interface conditions, an infinitely small fluid disk element ε is introduced with
small area A and thickness h, and with h A. The disk surfaces are parallel to the boundary
interface and one side of the surface is in the molecule dynamics region. Given a variation of the
disk position with δrε ,
(18)
.
= −δH FD
δrε
− δH FD/MD
δrε
i
αε
On the left-hand side
9
dt .
= ρAh du
i
αε
(19)
On the right-hand side, we first introduce the local coordinate system, which consists of one
normal direction (n) and two tangential directions (t,τ) at a certain point on the interface, i.e.,
n = cosα1i + cosα2j+ cosα3k
t = cosβ1i + cosβ2j+ cosβ3k
τ= cosγ1i + cosγ2j+ cosγ3k
.
⎧
⎪
⎨
⎪
⎩
(20)
only exerts on the disk surface in the fluid region, so that
Stress −δHFD
δrε
=σij inA = (− p + 2µ∂u
−
δHFD
δrε
⎡
⎢
⎣
∂nin)n + µ(∂(u⋅n)
∂t
+ ∂(u⋅t)
∂n
)t + µ(∂(u⋅n)
∂τ
+ ∂(u⋅τ)
∂n
⎤
⎥ A. (21)
)τ
⎦
The last term of eqn (18) can be worked out as
= −δUele
δrε
−δHMD/FD
δrε
2σpol DiiDo
1
Don
where − ∂Uele
∂rε
term − ∂Uhse
∂rε
= fdielecA =
nA is the dielectric boundary electrostatic force [87]. The
= −γκnA is the pressure and surface tension from the hard sphere entropy, aka the
− δUhse
δrε
− δUvdw
δrε
, (22)
hydrophobic term [84], where κ is the curvature. The van del Waals force, δUvdw
δrε
[84],
proportional to the volume of element Ah , can be ignored when comparing to the electrostatic
forces and surface tension as h is infinitely small. Combining eqns (18) and (19) and the terms
calculated above, the interface conditions can be summarized as
10
−p + pg −γκ+ fdielec + 2µ∂u
∂n in = 0
+ ∂(u⋅t)
∂n = 0
+ ∂(u⋅τ)
= 0,
∂n
∂t
∂(u⋅n)
∂τ
∂(u⋅n)
on ∂Ω (23)
3. Numerical Algorithms
We explored to implement the multi-scale model in numerical simulations with a strategy similar
to those of the classical Car-Parrinello molecular dynamics (CPMD) model [90], which can be
regarded as a multi-scale model via coupling equations of motion for ions and electrons in two
different mechanics. In CPMD electrons are treated as active degree of freedom, via fictitious
dynamics variable, and the fictitious electron dynamics is coupled with ionic dynamics in the
Berendsen heat bath to approach the Born-Oppenheimer surface. The CPMD model results in a
conservative ionic dynamics that is extremely close to the Born-Oppenheimer surface.
Our approach is to couple equations of motion for solute atoms and continuum solvent.
The solvent part is also treated by the fictitious dynamics variable, and since our model is based
on a finite-difference method, it is the fluid element. The fictitious fluid dynamics is modeled by
the incompressible Navier-Stokes (NS) equation. The fictitious fluid dynamics model is coupled
with all-atom molecular dynamics model in the Berendsen heat bath to approach the surface
provided by all-atom MD simulations at a preset temperature. In doing so, the changes to the
existing molecular mechanics simulation engine can be kept at the minimal and there is a very
clear boundary between the FD and MD simulation routines, facilitating the development of the
new model into a viable simulation engine for future biomolecular applications.
11
3.1 FD time integration
Our previous work has addressed the mathematical issues in solving fluid dynamics equations
numerically [61, 62]. After setting the water density to unity, the velocity can be solved by the
second-order semi-implicit backward Euler method as
where
3uk+1 − 4uk + uk−1
2Δt
+ u⋅∇u
(
)k+1 = −∇pk+1 + µΔuk+1 + Fk+1,outside
3uk+1 − 4uk + uk−1
2Δt
= µΔuk+1,inside
pk+1 = 2pk − pk−1
)k+1 = 2 u⋅∇u
(
u⋅∇u
(
)k − u⋅∇u
(
)k−1.
The pressure is solved by:
Δpk+1 = −∇ i((uk+1 i ∇)uk+1) + ∇ iFk+1.
(24)
(25)
(26)
A new issue facing the application of the FD model to molecular simulation is the
presence of van der Waals force ( F ), which has a large gradient nearby the interface because it is
too close to the solute atom centers. The large gradient is almost always challenging to address
with a finite-difference type of method. In this study, we overcome the issue by introducing a
variable p' , where
∇ ′p = µΔu without computing the numerical gradient of the van der Waals potential.
Specifically given p'(k+1) = pk+1 + Γk+1 , the equivalent form of eqn (26) to be solved numerically
′p = p + Γ with ∇Γ = −F obtained analytically. Therefore, we can solve
(27)
as
Δp'(k+1) = −∇ i((uk+1 i ∇)uk+1)
Accordingly, eqn (24) is updated as
12
3uk+1 − 4uk + uk−1
2Δt
+ u⋅∇u
(
)k+1 = −∇p'(k+1) + µΔuk+1,outside
3uk+1 − 4uk + uk−1
2Δt
= µΔuk+1,inside
p'(k+1) = 2p'(k) − p'(k−1)
where p'(k+1) is taken as
Finally the interface boundary condition eqn (23) becomes
−p' + Γ + pg −γκ+ fdielec + 2µ∂u
∂nin = 0
+ ∂(u⋅t)
∂n = 0
+ ∂(u⋅τ)
= 0,
∂n
∂t
∂(u⋅n)
∂τ
∂(u⋅n)
(28)
(29)
(30)
At each time step, the p' is interpolated with the one-side least square fitting method.[91] Γ is
computed analytically for each interface point where the interface boundary condition eqn (30) is
enforced. When doing so, we can completely avoid finite-difference operations involving van del
Waals energy and forces.
As presented in our previous works, the remaining major mathematical challenge in
solving these coupled partial differential equations is the presence of the free boundary condition
eqn (23) that allows the solute-solvent interface to equilibrate according to our physical model.
To enforce the free boundary condition when solving pressure or velocity, we utilized the jump
conditions of un and pn as the augmented variables, respectively [61, 62, 92]. The considerations
of augmented variables lead to extra correction terms on the right-hand side in eqns (24) and (26).
After the correction, each velocity component solver is equivalent to a Helmholtz equation. Once
the velocity is updated, the pressure solver is simplified to a Poisson equation. In this
13
implementation, we utilized the MICCG numerical solver to solve these linear differential
equations [93-96].
When solving the linear systems, the fluid domain is contained in a rectangular box, the
conditions at the outer boundary of the rectangular R are
u x=xmin = 0, u x=xmax = 0
∂u
= 0
∂y y=ymin
u z=zmin = 0, u z=zmax = 0
= 0, ∂u
∂y y=ymax
on ∂R , (31)
p = 0,
which represents a pipe flow in the y direction. The use of the boundary condition allows the
mass conservation law to be preserved since the incompressible solvent fluid can go in and out of
the simulation box freely.
3.2 FD/MD interface update
Once the fluid velocity field is known, the next step is to use it to update the solute/solvent
interface. The equivalent step in the solute region is to update particle positions based on particle
velocities. Numerically we use the level set method based on the finite-difference method [97-
99]. In the level set method, a scalar function, i.e. the level set function, is used to represent the
moving interface implicitly. The interface is located where the level set function is zero (d=0), i.e.
the zero level set Γ(t) = {y : d(y,t) = 0}. Suppose that Γ(t) moves according to velocity v:
), where v is known after the fluid dynamics equations are solved. Given the
∂Γ(t) / ∂t = v Γ(t)
interface velocity, if we want the level set function (d) to satisfy Γ(t) = {y : d(y,t) = 0} after
updating, we can impose the following equation upon d(y,t) [97-99]
(
∂d
∂t + v ⋅∇d = 0
(32)
14
with the initial condition Γ(0) = {y : d(y,0) = 0}, i.e. the level set function initially set for the
initial configuration in our case. Here the level set function was initially set as a signed distance
function to the solvent accessible surface with a specified solvent probe.
3.3 Overview of the FD/MD numerical procedure
In our system, the atomic details for the solute region are preserved, and the solvent region is
modeled as in 3.1. To simulate the solute particle dynamics, a standard MD engine with the
leapfrog time integrator [100] coupled to a heat bath is used. The temperature coupling is
realized with the Berendsen thermostat, which has been widely used in molecular simulation
community [101]. Once the heat bath is specificed, the procedure of the FD/MD can be
summarized into the following steps:
1. Input and initialize system parameters for solute atoms
such as temperature, number of particles, time step, etc.
Initialize initial positions and velocities of all solute
atoms;
2. Initialize FD simulation box and grid points. Initialize
velocity and pressure of fluid elements;
3. Compute energy and forces from the potential function of
solute atoms;
4. Compute van del Waals forces and pressure between solute
atoms and fluid atoms;
5. Use the particle MD engine to update new velocities and
positions of solute atoms;
6. Use the FD engine to update new velocities and pressures
of fluid elements;
7. Update new FD/MD interface;
15
8. Repeat steps 3-7.
Dynamics variables, such as position, velocity, pressure, and level set function, are periodically
stored after step 7 as requested. These can be used as input to restart the FD/MD simulation as
needed.
4. Other Computational Details
For the FD simulations, physical parameters of water are set as those at 300K with viscosity
µ= 8.51×10−4Pa ⋅s,
density ρ=1.00 ×103kg/m3,
tension
and
hydrophobic
surface
γ = 8.94 ×10−2 kcal/mol⋅A2 , with the later optimized for biomolecules given the SAS molecular
surface definition in a previous work for the Amber force fields [84]. The water probe was set as
1.0 Å to set up the initial SAS surface. In the FD simulation programs, both water viscosity and
density are often set as 1.0 in the internal unit. Thus proper interface between FD and MD
simulation portions of our model require careful unit conversion. The details in deriving these
conversion factors are given in Appendix, and the actual conversion factors are listed in Table I.
Variable
Time (t)
Density (ρ)
Energy (E)
MD Unit
1 ps
1 kg
1 kcal/mol
FD Unit
85.1
1.00 ×1027
9.60 ×10−2
Table I. Conversion factors between FD and MD engines.
The FD/MD multi-scale simulation engine was developed in a revised Amber 16 release
[63-65]. The Amber ff14 force field is used to generate the topology files and the TIP3P water
model is used to model the water molecules. All atomic charges were set to be zero to focus on
the nonelectrostatic interactions in this study. The simulations were conducted with bonds
involving hydrogen constrained. Time step was set to be 0.002 ps for both fluid dynamics region
and molecule dynamics region. The temperature coupling constant is 0.2 ps in the Berendsen's
16
thermostat to couple the temperature of the MD region, which is set to be 5 K to study the
relaxation of the solute-solvent interface in this study.
Since the goal of the current development is to evaluate how well the MD/FD method
reproduce the solvent interface, the MD region are restrained to focus on the FD simulation.
Given that the external forces on the FD region are only van der Waals force and hydrophobic
force, and they should be balanced each other at equilibrium. To speed up the relaxation, we
explored both artificially increase the external force terms (by a factor of 10) or decrease the
viscosity terms (by a factor of 10) to accelerate the relaxation towards equilibrium. It was found
that the low viscosity runs did not relax as fast as the high force runs (data not shown).
Nevertheless all alternatives will be further explored in a future study.
A single ion (Na+), a single molecule n-methyl amine (NMA), and four typical small
molecular complexes, adenine-thymine (AT), guanine-cytosine (GC), arginine-aspartic acid (RD)
and lysine-aspartic acid (KD) were chosen to analyze the solute-solvent surface produced by the
FD/MD method. In this stage of our development, the electrostatic interactions were turned off
so only van der Waals and hydrophobic interactions of the solute molecules were considered
though water molecules were not alternated. As a benchmark to evaluate the quality of the new
multi-scale model, we conducted all-atom molecular mechanics simulations for the four tested
dimer complexes to sample the solvent interface with explicit TIP3P water molecules. In these
simulations, all molecules first underwent a 10,000-step energy minimization starting with a
5,000-step steepest descent followed by a 5,000-step conjugate gradient minimization. Then all
solute atoms were restrained with a harmonic force constant of 50 kcal/mol-Å in all subsequent
heating, equilibration, and production simulations. The molecular dynamics simulations were
first heat up from 0K to 300K in 20 ps. This was then followed with a 10 ns simulation at the
17
constant temperature of 300K and the constant pressure of 1 bar with the Berendsen heat and
pressure baths. The water molecules sampled in the last 5 ns was used to analyze the solute-
solvent surfaces.
5. Results and Discussion
5.1 Single atom relaxation: Reproduction of analytical solution
We first validated the FD/MD engine with a simple system with analytical solution: the solute-
solvent interface of a single atom, given that the balance of hydrophobic force and van del Waals
force would lead to a final equilibrium surface, a sphere with radius of r0 . The equilibrium can
be analytically solved once the solvation free energy of the system is given as
G = ρ
+∞
A
⎛
∫
r12 −
⎝⎜
r0
A
⎛
= 4πρ
9r9 −
⎝⎜
B
⎞
⎠⎟ 4πr2 dr
r6
B
⎞
3r3 +γr2
⎠⎟
+γ4πr2
(33)
Starting from a given initial state, it is expected that the system converge to its free energy
minimum if there is no energy barrier, which is the case here.
In this test, an Amber sodium ion solvated in TIP3P water was used as an illustration.
With the specified surface tension and van der Waals parameters from Amber 14 force field, the
r4 + 2γr)
(34)
B
gradient of the free energy can be expressed as
A
r10 +
∂G
∂r = 4πρ(−
of
A = 4127 kcal/mol⋅A12,
the
values
and
Given
γ = 8.94 ×10−2 kcal/mol⋅A2, the numerical solution shows that there is only one root for
∂G
∂r = 0 when r is positive, which gives the radius of the sphere to be 2.45 Å. It is also clear that
B = 3.570 kcal/mol⋅A6,
18
∂r < 0 when r approaches 0+ and ∂G
∂G
∂r > 0 when r approaches infinity. Given that 1) the gradient
changes from negative to positive as r changes from 0+ to +∞, and 2) there is only one root for
the gradient, it can be concluded that the gradient is negative when r < 2.45 Å, and positive
when r > 2.45 Å. Thus free energy G is monotonically decreasing when r < 2.45 Å, and
monotonically increasing when r > 2.45 Å. This analysis shows that there is no energy barrier in
the physically allowed range of r.
Therefore it is possible to use a simple steepest descent minimization or a low-
temperature MD relaxation to reach the global minimum in the solvation free energy. Figure I
plots the evolution of volume versus time for the tested low-temperature relaxation run. It is
apparent that the volume of the solute-solvent interface quickly converges to a constant volume,
consistent with our analysis above. The numerical volume agrees with the analytical solution
with an error of ~0.3%. Note also that the equilibrium volume is a spherical sphere for the single
ion as expected.
19
Figure I. (1) Time evolution of volume (Å3) in the restrained FD/MD simulation of sodium ion.
(2) Spherical contour of solute-solvent interface when reaching the equilibrium.
5.2 Monomer relaxation: Symmetric interface
Next, we performed the low-temperature relaxation of NMA, a mirror-symmetrical monomer. As
shown in Figure II, the volume reaches the equilibrium value within 500 steps (1.0 ps). The
contour plot shows the symmetrical monomer possesses a symmetrical interface at equilibrium.
VMD visualization in 3D indicates that a detailed surface contour similar to that of the solvent
excluded surface can be found (see supplementary materials).
20
Figure II. (1) Time evolution of volume (Å3) in the restrained FD/MD simulation of NMA. (2)
SES-like solute-solvent interface is observed when reaching the equilibrium.
5.3 Dimer relaxation
Four typical small molecular complexes, adenine-thymine (AT), guanine-cytosine (GC),
arginine-aspartic acid (RD) and lysine-aspartic acid (KD) were tested to evaluate the
performance of the FD/MD simulation method. As shown in Figure III, the solute volumes reach
the equilibrium values within 500 steps (1.0 ps) for all four dimers. Figure IV presents the time
evolutions of force balancing on the solute-solvent interface. It is clear that the numerical solvent
pressure and viscosity pressure decrease significantly and approach zero as time goes on. On the
other hand, the hydrophobic (surface tension) pressure and the analytical van der Waals pressure
become the dominant components, reaching steady values while balancing each other out. This is
another evidence that the system approach equilibrium. Apparently the balance between
hydrophobic and van der Waals components is not perfect, due to the presence of residual fluid
flow nearby the solute. This issue will be addressed in our future refinement of the numerical
algorithm to be discussed below.
21
AT
GC
RD
KD
Figure III. Time evolutions of volume (Å3) in the restrained FD/MD simulations of dimers:
adenine-thymine (AT), guanine-cytosine (GC), arginine-aspartic acid (RD) and lysine-aspartic
acid (KD).
22
AT
GC
RD
KD
Figure IV. Time evolutions of average absolute pressure components on the solute-solvent
interface: adenine-thymine (AT), guanine-cytosine (GC), arginine-aspartic acid (RD) and lysine-
aspartic acid (KD). (a) Numerical solvent pressure; (b) Hydrophobic pressure; (c) Viscosity
pressure; (d) Analytical van der Waals pressure.
5.4 Comparison with explicit solvent simulations
Finally, a key issue in the current development of the FD/MD model is to see whether the model
at least qualitatively agrees with explicit solvent MD simulations. Discrepancy is possible given
that no optimization has been attempted. Therefore, it is interesting to analyze the solute-solvent
interfaces as sampled by both the FD/MD model and the explicit solvent MD model.
23
This analysis was conducted in the following manner. The water molecules in explicit
solvent MD simulations were sampled every 5 ps over the course of a 5 ns production run for
each tested dimer with all solute atoms restrained in the initial position. A total of 1000
snapshots were collected for visualization. To facilitate visualization, water molecules beyond
3.0 Å distance from any solute atom were discarded. The water distribution maps were used as
references to assess the solute-solvent surface sampled by the FD/MD simulation method. Figure
V shows the distribution of water oxygen atoms and the FD/MD surface when viewed outside of
the solute-solvent surface, and Figure VI shows the distribution and surface when viewed inside
of the solute-solvent surface. Overall, the FD/MD surfaces match very well with the solute-
solvent boundaries as sampled in the explicit MD simulations for all four tested complexes. Note
too there are a few places of discrepancies, which indicate that the parameters used in the
FD/MD model needs to be optimized. VMD visualization in 3D further illustrates the agreement
presented here (see supplementary materials).
24
Figure V. FD/MD surfaces (white wireframe) and water molecules from explicit MD
simulations (yellow dots) of four tested dimers: adenine-thymine (AT), guanine-cytosine (GC),
arginine-aspartic acid (RD) and lysine-aspartic acid (KD). Here viewer stands outside of the
surfaces.
25
Figure VI. FD/MD surfaces (white wireframe) and water molecules from explicit MD
simulations (yellow dots) of four tested dimers: adenine-thymine (AT), guanine-cytosine (GC),
arginine-aspartic acid (RD) and lysine-aspartic acid (KD). Here viewer stands inside of the
surfaces.
5.5 Limitations of the model and future directions
There are clearly limitations in the proposed FD/MD model. The first limitation is that we
artificially make both hydrophobic and van der Waals term 10 times higher to accelerate the
solute-solvent interface relaxation because the focus of the current model is for equilibrium
properties of the solute, but not the physically correct solvation relaxation process, which may be
26
important if the FD/MD model is applied to study hydrodynamic properties due to the presence
of the molecular solute. Nevertheless, the artificial setting does not affect the converged solute-
solvent interface because both hydrophobic and van der Waals pressure are simultaneously
increased. Secondly, the finite-difference grid spacing used in the FD engine is 0.5 Å, which is
widely used in biomolecular applications of a finite-difference method given a high enough
resolution of molecular surface topology can be achieved. However, the relatively fine grid also
leads to highly inefficient numerical procedure. To date we have not paid special attention to the
numerical efficiency of our FD engine, and this will be a focus in our future development. The
development and illustrations here mainly show that the FD/MD model is sound and it does
produce physically meaningful observations consistent with the all-atom MD model, which is
very promising.
As we pointed out in Other Computational Details the FD parameters for the water
solvation process was from a previous study to optimize a related nonpolar solvent model.
Apparently this is not optimal for the current FD/MD model. Our next step will be to investigate
how to optimize the hydrophobic term and van der Waals term to best reproduce the all-atom
explicit solvent MD model. In addition, we will also incorporate the electrostatic interaction as
modeled by the Poisson-Boltzmann method to build a more realistic FD/MD model for
biomolecular applications. To study fluid dynamic properties due to the presence of molecular
solutes, we think the best strategy is to incorporate a coarse-grained molecular model instead of
the all-atom model to make it a viable approach for systems with interesting hydrodynamic
properties.
6. Conclusions
27
A multi-scale framework was recently proposed for more realistic molecular dynamics
simulations in continuum solvent models by coupling a molecular mechanics treatment of solute
with a fluid mechanics treatment of solvent [61]. Our previous work addressed the mathematical
issues in solving fluid dynamics equations numerically [61, 62]. In this study we incorporated the
fluid dynamics algorithm with the Amber molecular mechanics package [63-65] to conduct
atomistic simulations of biomolecules. A major issue in the application of the multi-scale model
in atomistic simulations is the presence of van del Waals potential, which has a large gradient
nearby the solute-solvent interface. It is virtually impossible to treat van der Waals potentials
with any reasonably fine finite-difference method. We overcame the challenge by removing the
van del Waals potential from pressure when solving the finite-difference fluid dynamics
equations, and adding back the van del Waals potential analytically in the free-boundary
condition.
We first validated the FD/MD engine with a simple system with analytical solution: the
solute-solvent interface of a single atom. The balance of hydrophobic force and van del Waals
force would lead to the final equilibrium surface of a sphere. Our test shows that the volume of
the solute-solvent interface quickly converges to the analytical value with an error ~0.3%. Next,
we performed the relaxation of NMA, a mirror-symmetrical monomer. The contour plot shows
the symmetrical monomer possesses a symmetrical interface at equilibration. VMD visualization
in 3D indicates that a detailed surface contour similar to that of the solvent excluded surface can
be found. Four typical small molecular complexes were then tested to evaluate the performance
of the FD/MD simulation method. The solute volumes reach the equilibrium values within 1.0 ps
for all four dimers. The time evolutions of force balancing analysis on the solute-solvent
interface show that the numerical solvent pressure and viscosity pressure decrease significantly
28
and approach zero as simulation time goes on. On the other hand, the hydrophobic (surface
tension) pressure and the analytical van der Waals pressure become the dominant components,
reaching steady values while balancing each other out. This strongly indicates that the systems
approach the equilibrium at the end of the simulations.
Finally, a key issue at the current stage of the development is to investigate whether the
model at least qualitatively agrees with explicit solvent MD simulations. Therefore, it is
interesting to analyze the solute-solvent interfaces as sampled by both the FD/MD model and the
explicit solvent MD model. Comparisons show that the FD/MD surfaces agree very well with the
solute-solvent boundaries as sampled in the explicit MD simulations for all four tested dimers.
Note too a few places of discrepancies do exist, which indicate that the parameters used in the
FD/MD model needs to be optimized further to achieve higher consistency with the all-atom
explicit solvent MD model.
In our next phase of the development, we will further improve the quality of hydrophobic
and van der Waals terms of the FD model to best reproduce all-atom force field model. It is also
interesting to investigate the effect of incorporating the electrostatic forces into the FD/MD
model to evaluate its impact on both numerical stability and consistency for a range of model
systems. Finally it is also interesting to explore more efficient and more robust numerical FD
engines for routine applications to biomolecular systems.
Supplementary Materials
Supplementary materials are available online for the 3D visualization in VMD for monomer
NMA, and the four molecular dimer complexes: adenine-thymine (AT), guanine-cytosine (GC),
arginine-aspartic acid (RD) and lysine-aspartic acid (KD).
Acknowledgements
29
This work was supported in part by NIH/NIGMS (GM093040 & GM079383).
Appendix
Since FD programs can use any arbitrary length unit, we set 1 internal length unit as 1 Ångstrom.
Given the length unit settled and the water density (ρ=1.00 ×103kg/m3 ) set as 1 internal density
unit, the internal mass unit can be computed to be equivalent to 1.00 ×10−27kg. Next we can
utilize viscosity to compute the time conversion factor. Given the unit of viscosity
Pa ⋅s=kg/(m ⋅s) we can use the mass conversion factor to derive the time conversion factor as
follows
1 internal viscosity unit = 8.509 ×10−4kg/(m is)
= 8.509 ×10−4 ×1.00 ×1027 /(1010 ×T)
A.1
This leads to T = 8.51×1013, which means 1 s = 8.51×1013 internal time unit. And thus we have
1 ps =10−12s = 85.1 internal time unit. The energy unit of 1 kcal/mol can be converted as
1 kcal/mol = 6.948 ×10−21 J = 6.948 ×10−21 kg m2/s2
= 6.948 ×10−21 ×1027 ×(1010)2 /(8.509 ×1013)2internal energy unit
= 9.60 ×10−2 internal energy unit
A.2
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
M. E. Davis, and J. A. Mccammon, Chem Rev 90 (1990).
B. Honig, K. Sharp, and A. S. Yang, J. Phys. Chem. 97 (1993).
B. Honig, and A. Nicholls, Science 268 (1995).
D. Beglov, and B. Roux, Journal of Chemical Physics 104 (1996).
C. J. Cramer, and D. G. Truhlar, Chem Rev 99 (1999).
D. Bashford, and D. A. Case, Annual Review Of Physical Chemistry 51 (2000).
N. A. Baker, Curr. Opin. Struct. Biol. 15 (2005).
J. H. Chen, W. P. Im, and C. L. Brooks, Journal of the American Chemical Society 128 (2006).
30
[9]
M. Feig, J. Chocholousova, and S. Tanizaki, Theoretical Chemistry Accounts 116 (2006).
[10]
P. Koehl, Curr. Opin. Struct. Biol. 16 (2006).
[11] W. Im, J. H. Chen, and C. L. Brooks, Peptide Solvation and H-Bonds 72 (2006).
[12]
B. Z. Lu et al., Communications in Computational Physics 3 (2008).
[13]
J. Wang et al., Communications in Computational Physics 3 (2008).
[14] M. D. Altman et al., Journal of Computational Chemistry 30 (2009).
[15] Q. Cai et al., in Annual Reports in Computational Chemistry, edited by A. W. Ralph (Elsevier,
2012), pp. 149.
[16]
L. Xiao, C. Wang, and R. Luo, Journal of Theoretical and Computational Chemistry 13 (2014).
[17] W. M. Botello-Smith, Q. Cai, and R. Luo, Journal of Theoretical and Computational Chemistry
13 (2014).
[18]
[19]
R. E. Georgescu, E. G. Alexov, and M. R. Gunner, Biophysical Journal 83 (2002).
J. E. Nielsen, and J. A. McCammon, Protein Sci 12 (2003).
[20]
J. Warwicker, Protein Sci 13 (2004).
[21]
C. L. Tang et al., J. Mol. Biol. 366 (2007).
[22]
[23]
J. M. J. Swanson, R. H. Henchman, and J. A. McCammon, Biophysical Journal 86 (2004).
C. Bertonati, B. Honig, and E. Alexov, Biophysical Journal 92 (2007).
[24] A. R. Brice, and B. N. Dominy, Journal of Computational Chemistry 32 (2011).
[25]
[26]
R. Luo et al., Biophysical Journal 80 (2001).
L. David et al., Journal of Physical Chemistry B 103 (1999).
[27] D. Shivakumar, Y. Q. Deng, and B. Roux, J Chem Theory Comput 5 (2009).
[28] A. Nicholls et al., Journal of Medicinal Chemistry 51 (2008).
[29]
T. P. Korman et al., Biochemistry 47 (2008).
[30]
R. Luo et al., Biophysical Chemistry 78 (1999).
[31] K. L. Mardis, R. Luo, and M. K. Gilson, J. Mol. Biol. 309 (2001).
[32]
S. A. Marshall, C. L. Vizcarra, and S. L. Mayo, Protein Sci 14 (2005).
[33] M. J. Hsieh, and R. Luo, Proteins-Structure Function and Bioinformatics 56 (2004).
31
[34]
E. Z. Wen, and R. Luo, Journal of Chemical Physics 121 (2004).
[35]
[36]
E. Z. Wen et al., Journal of Molecular Graphics & Modelling 22 (2004).
T. Z. Lwin, and R. Luo, Journal of Chemical Physics 123, 194904 (2005).
[37]
T. Z. Lwin, R. H. Zhou, and R. Luo, Journal of Chemical Physics 124, 034902 (2006).
[38]
T. Z. Lwin, and R. Luo, Protein Science 15 (2006).
[39] Y.-H. Tan, and R. Luo, Journal of Physical Chemistry B 112 (2008).
[40] Y. Tan, and R. Luo, BMC Biophysics 2 (2009).
[41] Q. Lu, Y.-H. Tan, and R. Luo, Journal of Physical Chemistry B 111 (2007).
[42]
[43]
J. Wang et al., Biophysical Journal 95 (2008).
J. Warwicker, and H. C. Watson, J Mol Biol 157 (1982).
[44] D. Bashford, and M. Karplus, Biochemistry 29 (1990).
[45] A. Jeancharles et al., Journal of the American Chemical Society 113 (1991).
[46] M. K. Gilson, Curr. Opin. Struct. Biol. 5 (1995).
[47]
S. R. Edinger et al., J Phys Chem B 101 (1997).
[48]
C. Tan, L. Yang, and R. Luo, Journal of Physical Chemistry B 110 (2006).
[49] Q. Cai et al., Journal of Chemical Physics 130, 145101 (2009).
[50]
J. Wang et al., Chemical Physics Letters 468 (2009).
[51] X. Ye et al., Biophysical Journal 97 (2009).
[52] X. Ye, J. Wang, and R. Luo, Journal of Chemical Theory and Computation 6 (2010).
[53]
R. Luo, J. Moult, and M. K. Gilson, Journal of Physical Chemistry B 101 (1997).
[54]
J. Wang et al., Physical Chemistry Chemical Physics 12 (2010).
[55] M. J. Hsieh, and R. Luo, Journal of Molecular Modeling 17 (2011).
[56] Q. Cai et al., Journal of Chemical Theory and Computation 7 (2011).
[57]
J. Wang et al., Journal of Chemical Theory and Computation 8 (2012).
[58] X. Liu et al., Physical Chemistry Chemical Physics (2013).
[59]
C. Wang et al., Computational and Theoretical Chemistry 1024 (2013).
32
[60]
L. Xiao et al., The Journal of Chemical Physics 139 (2013).
[61]
[62]
L. Xiao et al., Chem Phys Lett 616-617 (2014).
Z. L. Li et al., Journal of Computational Physics 297 (2015).
[63] D. A. Case et al., (University of California, San Francisco, 2016).
[64] D. A. Case et al., (University of California, San Francisco, 2016).
[65] D. A. Case et al., Journal of Computational Chemistry 26 (2005).
[66]
P. K. Weiner, and P. A. Kollman, Journal of Computational Chemistry 2 (1981).
[67]
S. J. Weiner et al., Journal of the American Chemical Society 106 (1984).
[68] W. D. Cornell et al., Journal of the American Chemical Society 117 (1995).
[69]
J. M. Wang, P. Cieplak, and P. A. Kollman, Journal of Computational Chemistry 21 (2000).
[70] Y. Duan et al., Journal of Computational Chemistry 24 (2003).
[71]
[72]
L. J. Yang et al., Journal of Physical Chemistry B 110 (2006).
B. R. Brooks et al., Journal of Computational Chemistry 4 (1983).
[73] A. D. MacKerell et al., J Phys Chem B 102 (1998).
[74] N. Foloppe, and A. D. J. MacKerell, J. Comp. Chem. 21 (2000).
[75] W. L. Jorgensen, and J. Tiradorives, Journal of the American Chemical Society 110 (1988).
[76] W. L. Jorgensen, D. S. Maxwell, and J. TiradoRives, Journal of the American Chemical Society
118 (1996).
[77] G. A. Kaminski et al., J Phys Chem B 105 (2001).
[78]
E. S. Reiner, and C. J. Radke, J Chem Soc Faraday T 86 (1990).
[79] K. A. Sharp, and B. Honig, J. Phys. Chem. 94 (1990).
[80]
[81]
L. Xiao et al., Journal of Chemical Physics 139 (2013).
L. Xiao, C. H. Wang, and R. Luo, J Theor Comput Chem 13 (2014).
[82] D. Chandler, Nature 437 (2005).
[83]
J. Dzubiella, J. M. J. Swanson, and J. A. McCammon, Phys Rev Lett 96 (2006).
[84]
[85]
C. Tan, Y. H. Tan, and R. Luo, J Phys Chem B 111 (2007).
L. T. Cheng et al., Journal of Chemical Physics 127 (2007).
33
[86]
P. W. Bates et al., J Math Biol 59 (2009).
[87]
L. Xiao et al., The Journal of chemical physics 139 (2013).
[88]
B. D.Jones, (Submitted 2014).
J. D. W. Alexander L.Fetter, Theoretical Mechanics of Particles and Continua (Dover
[89]
Publications Inc., Mineola, New York, 2003).
[90]
[91]
R. Car, and M. Parrinello, Phys Rev Lett 55 (1985).
J. Wang et al., Chemical Physics Letters 468 (2009).
[92]
Z. L. Li et al., Comput Fluids 39 (2010).
[93]
J. Wang, and R. Luo, Journal of Computational Chemistry 31 (2010).
[94] Q. Cai et al., Journal of Chemical Theory and Computation 6 (2010).
[95]
R. Luo, L. David, and M. K. Gilson, Journal of Computational Chemistry 23 (2002).
[96] Q. Lu, and R. Luo, Journal of Chemical Physics 119 (2003).
[97]
S. Osher, and J. Sethian, Journal of Computational Physics 79 (1988).
J. A. Sethian, Level set methods and fast marching methods (Cambridge University Press,
[98]
Cambridge, 1999).
S. Osher, and R. Fedkiw, Level Set Methods and Dynamic Implicit Surfaces (Springer, Berlin,
[99]
2002).
[100] R. W. H. J. W. EastWood, Computer Simulations using Particles (McGraw-Hill, 1981).
[101] H. J. C. a. V. G. Berendsen, W.F., (1984).
34
|
1601.02770 | 1 | 1601 | 2016-01-12T08:54:51 | Inversion of hematocrit partition at microfluidic bifurcations | [
"physics.bio-ph",
"cond-mat.soft",
"physics.flu-dyn"
] | Partitioning of red blood cells (RBCs) at the level of bifurcations in the microcirculatory system affects many physiological functions yet it remains poorly understood. We address this problem by using T-shaped microfluidic bifurcations as a model. Our computer simulations and in vitro experiments reveal that the hematocrit ($\phi_0$) partition depends strongly on RBC deformability, as long as $\phi_0 <20$% (within the normal range in microcirculation), and can even lead to complete deprivation of RBCs in a child branch. Furthermore, we discover a deviation from the Zweifach-Fung effect which states that the child branch with lower flow rate recruits less RBCs than the higher flow rate child branch. At small enough $\phi_0$, we get the inverse scenario, and the hematocrit in the lower flow rate child branch is even higher than in the parent vessel. We explain this result by an intricate up-stream RBC organization and we highlight the extreme dependence of RBC transport on geometrical and cell mechanical properties. These parameters can lead to unexpected behaviors with consequences on the microcirculatory function and oxygen delivery in healthy and pathological conditions. | physics.bio-ph | physics |
Inversion of hematocrit partition at microfluidic
bifurcations
Zaiyi Shen,1 Gwennou Coupier,1 Badr Kaoui,2,3 Benoıt Polack,4,5
Jens Harting,6,7 Chaouqi Misbah,1 Thomas Podgorski1∗
January 13, 2016
affiliations: 1 Laboratoire Interdisciplinaire de Physique (LIPhy) UMR5588 CNRS-Universit´e
Grenoble Alpes, Grenoble F-38041, France ; 2 CNRS-Sorbonne University, Universit´e de
Technologie de Compi`egne, UMR7338 Biomechanics and Bioengineering, 60203 Compi`egne,
France; 3 TP1, University of Bayreuth, D-95447 Bayreuth, Germany; 4 Laboratoire d’H´ematologie,
CHU, Grenoble, France; 5TIMC-IMAG/TheREx, CNRS UMR5525, Universit´e Grenoble
Alpes, Grenoble, France; 6 Department of Applied Physics, Eindhoven University of Tech-
nology, P.O. Box 513, 5600MB Eindhoven, The Netherlands; 7 Faculty of Science and Tech-
nology, Mesa+ Institute, University of Twente, 7500 AE Enschede, The Netherlands
abbreviated title: Inversion of hematocrit partition
corresponding author: Thomas Podgorski, e-mail: [email protected]
Abstract
Partitioning of red blood cells (RBCs) at the level of bifurcations in the microcirculatory
system affects many physiological functions yet it remains poorly understood. We address this
problem by using T-shaped microfluidic bifurcations as a model. Our computer simulations
and in vitro experiments reveal that the hematocrit (φ0) partition depends strongly on RBC de-
formability, as long as φ0 < 20% (within the normal range in microcirculation), and can even
lead to complete deprivation of RBCs in a child branch. Furthermore, we discover a deviation
from the Zweifach-Fung effect which states that the child branch with lower flow rate recruits
less RBCs than the higher flow rate child branch. At small enough φ0, we get the inverse sce-
nario, and the hematocrit in the lower flow rate child branch is even higher than in the parent
vessel. We explain this result by an intricate up-stream RBC organization and we highlight the
extreme dependence of RBC transport on geometrical and cell mechanical properties. These
parameters can lead to unexpected behaviors with consequences on the microcirculatory func-
tion and oxygen delivery in healthy and pathological conditions.
keywords: microcirculation; blood; red blood cell; microfluidics; lattice Boltzmann method
1
1
Introduction
Blood flows through a complex network of the circulatory system – from large arteries to very tiny
capillaries – in order to ensure oxygen delivery and to remove metabolic waste. This task is mainly
carried out by red blood cells (RBCs) that are remarkably deformable, in healthy conditions, and
therefore able to squeeze into tiny capillaries. A change in rheological and flow properties of
the blood is often associated with hematological diseases or disorders (Fedosov et al., 2011) (e.g.
sickle-cell anemia, malaria, polycythemia vera). Understanding blood flow and its dependence on
the mechanical properties of its constituents may improve and lead to new applications in biomed-
ical technology, for example in blood substitutes development and transfusion techniques.
A major open problem in blood circulation is to understand the perfusion in the vasculature
networks, especially in the microvasculature where RBCs accomplish their vital functions. For
example, an improper hematocrit distribution is observed in heart microcirculation with conse-
quences such as occlusion zones (within many patients with apparently healthy coronary arteries).
These abnormal traffic zones cause a lack of oxygen supply to tissues that leads to cardiac is-
chemic disease (Cokkinos et al., 2006). The precise origin of this dysfunction is still a matter of
debate. The principal mechanism that dictates blood heterogeneity in the microvascular networks
is the hematocrit partition at the level of bifurcations. RBCs do not behave as passive tracers. Their
shape flexibility and dynamics have a decisive role because their size is comparable to that of blood
capillaries. A well known phenomenon in microcirculation is the Zweifach-Fung effect (Dellimore
et al., 1983; Fenton et al., 1985; Guibert et al., 2010; Pries et al., 1989): If we consider a bifurcation
(as in Figure 1), the child branch with the lower flow rate is depleted in RBCs as compared to the
parent vessel, while the other, higher flow rate child branch is enriched. That is, if in the parent
vessel the total volumetric flow rate is Q0 and the RBC volumetric flux is N0, and in the child
branch with the lower flow rate this flow rate is Q1 and the RBC flux N1, then N1/N0 < Q1/Q0.
When the flow rate is sufficiently small, the hematocrit in the child branch can even drop down to
zero, while it reaches high values in the other branch. Obviously, this phenomenon has physio-
logical consequences as it alters the transport of oxygen and other essential metabolites, and may
even trigger pathological disorders (e.g. occlusions in high hematocrit regions where the viscosity
is higher and cell adhesion is favored).
Previous studies have investigated the impact of the confinement (Barber et al., 2008; Chien
et al., 1985; Doyeux et al., 2011; Fenton et al., 1985), the bifurcation geometry (Audet and Ol-
bricht, 1987; Hyakutake and Nagai, 2015; Roberts and Olbricht, 2003, 2006; Woolfenden and
Blyth, 2011), the hematocrit (Ditchfield and Olbricht, 1996; Fenton et al., 1985; Roberts and Ol-
bricht, 2003; Yin et al., 2013), and the RBCs deformability (Barber et al., 2008; Li et al., 2012;
Xiong and Zhang, 2012; Yin et al., 2013) and aggregation (Sherwood et al., 2012; Yin et al., 2013).
Most of these parameters influence RBCs distribution in the feeding flow, which is believed to dic-
tates the partition at the bifurcation (Doyeux et al., 2011; Fenton et al., 1985; Li et al., 2012; Yin
et al., 2013). The Zweifach-Fung effect results from the existence of a cell free layer (CFL) close
to the walls, which is only occupied by plasma. The feeding flow is divided by a separating stream-
line into two parts, one feeding the low flow rate branch and the other feeding the high flow rate
branch. Due to the CFL, the RBC fraction entering the low flow rate branch is smaller compared
to the original RBC fraction in the total feeding flow. The depletion in the low flow rate branch
is accompanied by enrichment in the high flow rate branch. In addition to the CFL as the main
cause of the Zweifach-Fung effect, it has been argued that there is a relatively small counteracting
2
Figure 1: Snapshots of the RBCs partition, in both experiments and simulations, when the hema-
tocrit of the feeding flow is around φ0 =10%. The length ratio between the two child branches is
set to 3. A, B: Low viscosity contrast (experiments with λ =0.85 and simulation with λ =1). C, D:
High viscosity contrast (experiments with λ =10.3 and simulation with λ =10).
effect where cells entering the bifurcation tend to be displaced towards the low flow rate branch
compared to fluid streamlines (a fact that slightly reduces the Zweifach-Fung effect) (Barber et al.,
2008; Doyeux et al., 2011; Li et al., 2012; Ollila et al., 2013), but this question is still debated
(Hyakutake and Nagai, 2015; Xiong and Zhang, 2012).
The existence of a CFL is a consequence of the lateral migration of RBCs towards the vessel
center. This migration is a result of the wall-induced lift force due to hydrodynamic interactions
(Callens et al., 2008; Cantat and Misbah, 1999; Grandchamp et al., 2013), that depends on the
nature of RBC dynamics (like tank-treading or tumbling (Abkarian et al., 2007; Dupire et al.,
2012)). The final distribution of RBCs in a confined straight vessel is not only due to the lift force,
but it is also influenced by other factors: (i) The curved velocity profile of the Poiseuille flow
(Coupier et al., 2008; Farutin and Misbah, 2014; Katanov et al., 2015; Shi et al., 2012), and (ii)
The cell-cell hydrodynamic interactions (Grandchamp et al., 2013; Hariprasad and Secomb, 2014;
Katanov et al., 2015; Kruger et al., 2014; McWhirter et al., 2009).
In the present work, we study the hematocrit partition at bifurcations using two-dimensional
lattice Boltzmann simulations, whose outcomes are validated and supported by microfluidic exper-
iments. We show that RBCs deformability strongly impacts partition as long as the hematocrit is
below 20% (within the normal range in microcirculation). RBC deformability is governed by sev-
eral parameters such as membrane stiffness (shear, dilatation and bending elastic moduli), swelling
degree, membrane viscosity and the viscosity contrast between the hemoglobin and the suspend-
3
ϕ0=10%ALowviscositycontrastHighviscositycontrastϕ1CDBLowviscositycontrastHighviscositycontrastL1L2InletOutletExperiment, λ=0.85 Experiment, λ=10.3 Simulation, λ=1Simulation, λ=10ing fluid. Here we choose to tune the deformability through the latter parameter, the viscosity
contrast, that controls the RBC dynamics (tank-treading, tumbling or swinging) then all the mi-
gration mechanisms at the origin of the CFL. On the other hand, and more importantly, this study
reveals that hematocrit partition can be completely reversed, that is the low flow rate child branch
can be enriched in RBCs compared to the parent vessel. This newly reported effect is an outcome
of a subtle RBCs structuration in the microcirculatory system. This highlights the importance of
the notion of RBCs spatio-temporal organization as the main non-negligible ingredient to further
understand blood perfusion in the microvasculature.
2 Materials and methods
Design of the microfluidic bifurcations
In both simulations and experiments, we use T-shaped bifurcations such as shown in Fig.1: A
parent channel divides into two child branches with the same width, but with different lengths
L1 and L2 (L1 > L2). The ratio of the flow rates in branches 1 and 2 is then given by Q1/Q2 =
(η2L2)/(η1L1), where η1 and η2 are the apparent viscosities of the suspension in branches 1 and 2,
respectively. For dilute suspensions, where the viscosity is close to that of the suspending fluid, we
simply have Q1/Q2 = L2/L1. In simulations, we set the width of the channels to W = 20µm and
we vary L1/L2 from 1.43 to 3. In experiments, we have L1/L2 = 3, W = 20µm and the height of
the channel h is 8 µm. The length of the parent vessel was chosen as long as possible to allow for
the development of a stationary distribution of RBCs across the channel in the feeding flow (5 mm
in experiments and 1.5 mm in simulations). Microfluidic channels were produced by standard soft
lithography techniques, with molded PDMS bonded to glass. The RBC suspensions were perfused
by a syringe pump (KDS Legato 180) and imaging was performed by a video camera (Imaging
Source DMK 31AF03) mounted on an inverted microscope with motorized stage (Olympus IX71)
and a blue filter (434± 25 nm) corresponding to an absorption peak of hemoglobin.
Blood preparation
Blood samples were provided by the Etablissement Franc¸ais du Sang (EFS Rhone-Alpes) from
healthy donors. RBCs were isolated by centrifugation after being washed twice in phosphate buffer
saline (PBS) supplemented by 0.1 % bovine serum albumin (BSA). To prevent sedimentation of
RBCs in channels, the RBCs were re-suspended in density matching PBS and BSA solutions in
a mixture of water and iodixanol (Optiprep from Axis-Shield). This iso-dense solution was used
either alone (1.94 mPa.s at 20 ◦C) or after adding 5% dextran of molecular weight 2×106 (viscosity
23.4 mPa.s at 20 ◦C). The viscosity of the internal hemoglobin solution of healthy RBCs is around
20 mPa.s at 20 ◦C (Kelemen et al., 2001). This provides two values of the viscosity contrast λ ,
namely 10.3 and 0.85. The first value corresponds to the blood at 20 ◦C, while the physiological
value at body temperature is around 5-6 (Cokelet and Meiselman, 1968).
Note that we chose to vary the viscosity contrast λ as one way to tune deformability, and there-
fore the dynamics of lift and hydrodynamic interactions of cells. Stiffening cells using diamide or
glutaraldehyde was another possibility. However, from the experimental viewpoint, working with
hardened cells at high volume fractions in such a confined environment is quite difficult due to jam-
4
ming. It would have been nearly impossible to inject a suspension of very stiff cells at hematocrits
larger than 10%. Also, the dynamics of glutaraldehyde hardened cells is pure tumbling, which cor-
responds to very high values of the viscosity ratio λ . We do not expect the dynamics (and therefore
phase separation) to change much at values of λ greater than 10 and we found more interesting to
increase deformability by decreasing λ rather than trying to investigate less deformable cells (with
the experimental difficulties mentioned above).
Hematocrit measurements
Local hematocrit measurements were made by comparing suspension flow images to a reference
image without RBC, under identical illumination, and using the Beer-Lambert law of absorption.
The absorption coefficient was determined by a calibration with images at low hematocrit, where
a direct measurement can be made by counting individual cells. Hematocrit in branch i will be
denoted Hi. In experiments, h is small enough (8µm) so that the flow is quasi two-dimensional.
To allow a qualitative comparison with 2D numerical simulations, an area hematocrit φi was also
derived by multiplying the number of cells per unit area by the average cross-sectional area of
RBCs (S =19.8 µm2). Hi and φi are therefore linked by the relationship Hi = φiv/(Sh), where
v = 90 µm3 is the average volume of one cell.
Simulation method
In simulations, we use lattice Boltzmann method (LBM) to compute the fluid flow (Kaoui et al.,
2011; Zhang et al., 2007). Each RBC is modeled by 60 nodes interconnected by a potential that
allows bending, as well as a stretching modulus that penalizes distance variations between two
adjacent nodes. This achieves the RBC membrane incompressibility (Tsubota et al., 2006). In other
words, we set the spring constant to values as large as possible in order to keep the ratio between
the membrane perimeter and area constant. We define the reduced area as 4πA/p2 (with p is the
perimeter and A the enclosed area), which we set to 0.7 to produce a RBCs with a biconcave shape.
We use the immersed boundary method (IBM) to couple the fluid flow and RBC deformation
(Kaoui et al., 2011; Peskin, 2002; Zhang et al., 2007). For comparison with experiments, we set
the viscosity contrast to λ = 1 and λ = 10.
3 Results
As a guideline, we shall refer to the empirical law of Pries et al. taken from refs (Pries et al., 1989,
1990) that gives the hematocrit partition at a bifurcation:
logit(cid:0)H1Q1
(cid:1) = αlogit(cid:0)Q1/Q0 − β
(cid:1),
H0Q0
1− 2β
(1)
where α = 1 + 6.98(1− H0F )/a, β = 0.4/a (with a the tube diameter in microns) and logit(x) =
ln[x/(1− x)]. H0 is the volumetric hematocrit in the parent feeding branch, while H1 is the hema-
tocrit in a child branch. H0F is the feeding hematocrit in a reservoir that would be located right
5
Figure 2: The relative hematocrit in the low flow rate branch as a function of the hematocrit in the
parent vessel. The length ratio between the branches L1/L2 is set to 3. A: Simulations with λ =1
and λ =10. B: Comparison between experiments (λ =0.85 and λ =10.3) and the empirical law of
Pries et al. (Pries et al., 1990) (solid line), for the same cross-sectional area. The non-monotonous
evolution of the relative hematocrit at low H0 and high λ is related to the inversion of the Zweifach-
Fung effect, on which we comment later on (see Fig. 7).
before the narrow feeding vessel. Due to the Fahraeus effect, H0F is larger than H0 and a relation-
ship between both quantities is also given in Pries et al. (1990):
H0/H0F = H0F + (1− H0F )(1 + 1.7e−0.415a − 0.6e−0.011a).
(2)
Note that the partition law (Eq. 1) has been validated through in-vivo experiments with rats (thus
at body temperature), with narrow capillaries (of diameters a lower than 30 microns), but with
feeding hematocrits higher than 20%.
As we shall compare predictions for 3D hematocrits in a cylindrical tube with either 2D simula-
tions or experiments in a rectangular channel, we should avoid any direct quantitative comparisons,
but rather use Pries et al. predictions as a guideline to identify where new behavior is exhibited.
For comparison with simulations, we set a = W , where W is the channel width, and we shall con-
sider only the hematocrit ratios. For the experiments in rectangular channels, we set a to adjust the
cross-sectional areas: πa2/4 = W h.
6
0.20.40.60.8100.10.20.3H1/H0H0λ=0.85λ=10.3Empirical,A.Priesetal.AB0.20.40.60.8100.10.20.30.4ϕ1/ϕ0ϕ0λ=1λ=10The role of interactions
We analyze in details how RBC deformability affects the hematocrit partition at the bifurcations.
Figure 1 illustrates the Zweifach-Fung effect, observed in both experiments and simulations, at a
feeding area hematocrit of 10%. In both cases, less RBCs enter the low flow rate branch (the long
branch) simply due to the flow rates differences between the two child branches. However, the
asymmetry is significantly pronounced at low viscosity contrast λ (when the suspending fluid is
more viscous than the hemoglobin). To quantify the partition asymmetry, we measure the relative
hematocrit φ1/φ0 (or, equivalently, H1/H0), in the low flow rate branch, while we vary the hemat-
ocrit in the parent branch (Fig. 2A). Either in the simulations (Fig. 2A) or in the experiments (Fig.
2B), we see less RBCs in the low flow rate branch than in the parent one (φ1/φ0 < 1), when the
inlet hematocrits (φ0 or H0) lies between 5% and 45%, which is precisely a manifestation of the
Zweifach-Fung effect. When the viscosity contrast is low, we observe a significantly strong reduc-
tion of hematocrit in the low flow rate branch, both in experiments and simulations, at moderate
inlet hematocrit. This interesting observation suggests that the RBCs mechanical properties can
strongly impact the hematocrit partition in-vivo since the normal hematocrit is usually less than
20% (typically between 10 and 20 % (Fung, 2013)) in microcirculation.
However, when the hematocrit is high enough, the viscosity contrast plays a minor role. This
is clear in simulations (for φ0 larger than 25%, Fig. 21). Similarly, in the experiments, above
H0 = 20%, data for both λ converge to the Pries et al. prediction. The insensitivity to the viscosity
contrast beyond a critical hematocrit (φ0 (cid:39) 25%) is a robust phenomenon that is independent of the
length ratio between the branches (i.e. roughly the bulk flow rate ratio), as illustrated on Figure 3).
It is appealing to suggest that the dependency of the hematocrit partition upon the feeding
hematocrit is the result of the up-stream organization of RBCs in the parent vessel due to hydrody-
namic interactions. At low hematocrit flows, the cell-cell interaction is weak and the organization
of RBCs, within the vessel, depends mainly on the dynamics of each RBCs, thus on λ . The RBCs
aggregate at the center of the vessel due to the wall-induced lift force, that increases with decreas-
ing λ and increasing RBCs deformability (Grandchamp et al., 2013). This means that suspensions
of RBCs with high viscosity contrasts have wider distributions (smaller CFL) in the channel as
compared to suspensions having lower viscosity contrast. As a consequence, the asymmetry in
the partition is expected to increase when the viscosity contrast decreases, as shown in Figure 3.
To support this argument, the CFL thickness in the parent vessel and the configuration of RBCs
before the bifurcation are reported in Figure 4A,B. We can clearly see that RBCs distribution at
high viscosity contrast (λ =10 in simulations and λ =10.3 in experiments) is wider than that at low
viscosity contrast (λ =1 in simulations and λ =0.85 in experiments) when the feeding hematocrit is
low (φ0 <20%, see also Figure 4C-I, II).
However, when the hematocrit increases, the contribution of the hydrodynamic interactions
among RBCs becomes stronger and stronger. This causes a broadening of the distribution that
acts against the lift force. Consequently, the partition between the two branches becomes more
symmetric (that is φ1/φ0 becomes close to 1). Interestingly, those broad distributions are quasi
independent of the viscosity contrast (see Figure 4A and C-III, IV). Consequently, φ1/φ0 does not
depend on λ either (see Figure 3).
Thus the distribution is independent of the strength of the interactions between cells and be-
tween cells and walls, but it is mainly caused by geometrical constraints. In other words, interaction
between cells and the lift forces both depend on λ , and this result indicates that they depend more
7
Figure 3: Simulations: the relative hematocrit in low flow rate branch φ1/φ0 as a function of the
hematocrit in the parent vessel φ0, for several branches length ratios L1/L2 and viscosity contrasts
λ . At low enough φ0, the asymmetry between the two daughter branches is strongly enhanced
as the viscosity contrast λ is decreased, while the partitioning becomes independent on λ for
hematocrits above 20%.
or less on λ in the same way. The λ contributions cancel out once a critical feeding hematocrit
is reached. Noteworthy, beyond this critical hematocrit the separating ratio φ1/φ0 quasi plateaus
which enforces the idea that in this regime, the feeding flow can be considered as a three-layer
fluid (fluid-cell-fluid). The width of each layer will depend neither on the strength of interaction
(which is related to deformability) nor on the volume fraction.
The low hematocrit limit: Inversion of the Zweifach-Fung effect
Now we focus on the low hematocrit case, for which the partition depends strongly on the detail of
the interactions and on the volume fraction. Moreover a peculiar effect arises due to the prevalence
of the discrete nature of blood at that scale. For all hematocrits, the distribution of RBCs is not
homogeneous, but rather exhibits two lateral peaks (Figure 4). This become more pronounced at
low hematocrit (φ0 (cid:46) 5%), where a two-file distribution of RBCs is observed, as shown in Figure
5. For λ =10, there is almost no cell flowing in the central part of the vessel, even though the
wall-lift force tends to center them. The structure adopted by the suspension can be viewed as a
juxtaposition of layers with high and low hematocrits. For example, as shown in Figure 5 (top
8
00.510.050.150.250.35ϕ1/ϕ0ϕ0L1/L2=1.43λ=1λ=1000.510.050.150.250.35ϕ1/ϕ0ϕ0L1/L2=1.86λ=1λ=1000.510.050.150.250.35ϕ1/ϕ0ϕ0L1/L2=2.29λ=1λ=1000.510.050.150.250.35ϕ1/ϕ0ϕ0L1/L2=2.71λ=1λ=10ABCDFigure 4: A: CFL thickness as a function of the hematocrit in the parent vessel, for λ =1 and λ =10,
in simulations. Insets : Snapshots showing the suspension at the bifurcation. We define the CFL as
a layer where the integrated concentration profile is below 5% (Kumar et al., 2014). B: Snapshots
from experiments, for λ =0.85 and 10.3, and hematocrits φ0 =10, 20 and 30%. Every snapshot is a
superimposition of 10 successive images in order to highlight the CFL in the parent vessel. C: The
stationary volume fraction density functions in the parent vessel obtained from simulations.
9
345670.050.10.150.20.250.30.350.4CFLthickness(µm)ϕ0λ=1λ=10ϕ010%20%30%λ=0.85λ=10.300.20.40.60.805101520ϕy(µm)ϕ0=10%λ=1λ=1000.20.40.60.805101520ϕy(µm)ϕ0=20%λ=1λ=1000.20.40.60.805101520ϕy(µm)ϕ0=30%λ=1λ=1000.20.40.60.805101520ϕy(µm)ϕ0=40%λ=1λ=10IIIIIIVIIIIIIIIVABCFigure 5: The profiles of the hematocrit distribution and the corresponding snapshots of RBCs
distribution in the parent vessel. The feeding hematocrits are 5% and 40%, for two different values
of λ .
panel with λ = 10) the central part is depleted in RBCs, but it is escorted by two enriched layers,
which themselves are surrounded by two depleted layers at the periphery (close to the channel
walls). This 5-layer configuration (fluid-cell-fluid-cell-fluid) has an extremely interesting impact
on the partition. This can be highlighted by measuring φ1/φ0 as a function of the bulk flow rate
ratio Q1/Q0 between a child branch and the parent vessel, for fixed φ0 (Figure 6).
If we focus first on the results for high φ0 (φ0 (cid:39) 40%), we find again the insensitivity to λ . As
Q1/Q0 is increased from 0 to 0.5, the low flow rate branch 1 recruits first the CFL and then the
cells. This implies that φ1 starts at 0 and increases until reaching φ0 when the situation is symmetric
(Q1 = Q2 = Q0/2). Our results agree with a previous 2D simulation obtained for φ0 (cid:39) 32% (Yin
et al., 2013) as well as with the empirical law of Pries and coworkers (Pries et al., 1990).
An unexpected phenomenon is observed at low enough hematocrit, for φ0 = 5% and at high
viscosity contrast λ = 10, in contrast to the high hematocrit regime (see Figure 6 B, λ = 10). At
low Q1/Q0, the peripheral CFL in the parent vessel is recruited by the branch 1 so φ1 starts at 0 and
increases when Q1 increases. Around Q1/Q0 = 0.3, φ1 becomes larger than φ0. This means the
hematocrit is increased in the low flow rate branch, which is the reverse behavior of the Zweifach-
Fung effect. The five-layer structure mentioned above is the key ingredient for understanding
this unexpected behavior: In the intermediate range 0.3 < Q1/Q0 < 0.5, the low flow rate branch
recruits the lateral CFL layer plus the adjacent RBC-rich layer among the five layers. By contrast,
the high flow rate branch recruits the CFL layer close to the opposite wall plus its adjacent RBC-
rich layer (exactly as the low flow rate branch) as well as the central (and depleted) layer. Thus,
while both branches recruit approximately the same amount of cells per unit time, those are more
dilute in the high flow rate branch, which receives more fluid, while in the classical Zweifach-Fung
effect, the high flow rate branch is the one that receives more cells. For λ = 1 (see Figure 6B with
λ = 1), the two-peak structure is not as marked as in the case of λ = 10, so the reverse Zweifach-
Fung effect is not as strong. The subtle role played by the suspension structuring at low hematocrit
is also supported by our experiments, where the interplay between the diffusion and the wall-lift
force is controlled by varying the width W of the inlet channel (see Figure 7). When W is low (W
10
0510152000.050.10.150.2y(µm)ϕϕ0=5%λ=1λ=100510152000.20.40.60.8y(µm)ϕϕ0=40%λ=1λ=10Figure 6: The relative hematocrit in one child branch as a function of the bulk flow ratio. Solid lines
correspond to the empirical law proposed in ref (Pries et al., 1990), for a = W . For simulations, the
relative hematocrit is φ1/φ0. For Pries law, it is given by H1/H0. A: high hematocrit (φ0 = H0 =
40%). B: low hematocrit (φ0 = H0 = 5%).
= 10 µm), the hematocrit in the low flow rate branch is much lower than expected from Pries et al.
predictions (which were not validated on this confinement range). This is caused by the CFL effect
that becomes very strong. As in the simulations, for Q1/Q0 = 0.25, this effect is more pronounced
at low λ , that corresponds to a more important wall lift force. The 5-layer structure is clearly
observed also in the experiments for H0 < 5%, W = 20 or 30 µm and at low λ (Figure 7C), but
not at high λ , while it was more strongly marked at high λ in the simulations. This indicates that
this peculiar structure is very dependent on the mechanical properties of the cells and also on the
degree of confinement. Nevertheless, a robust feature is valid in both simulations and experiments:
When the two-file structure of RBCs takes place, a clear inversion of the blood partition at the
bifurcation is observed. In the experiments, this corresponds to Figure 7A, where some points lie
above the equal partition line when H0 < 2%.
There are also situations in which one of the two branches can be even completely devoid of
RBCs (Figure 7C). A corresponding prolonged lack of RBCs perfusion to real blood vessels causes
dysfunction and possibly ischemia disease. Because RBC mechanical properties are affected by
aging and pathologies, these can induce abnormal partitions of the hematocrit in the vascular net-
work.
11
00.30.60.91.21.50.20.30.40.50.60.70.8Relative HctQ1/Q0Highhematocritλ=1,simulationλ=10,simulationEmpiricalRelative Hct00.30.60.91.21.50.20.30.40.50.60.70.8Q1/Q0Lowhematocritλ=1,simulationλ=10,simulationEmpiricalABFigure 7: Experiments: the hematocrit in the low flow rate branch H1 as a function of the hematocrit
in the parent vessel H0. A: Low λ ; B: High λ . The length ratio between the branches is set to
3. The width W of the inlet channel is set to 10, 20 and 30 µm. The corresponding empirical
laws of Pries et al.(Pries et al., 1990) are also shown. The dotted line (the one with highest slope)
corresponds to equal partition (H1 = H0). For W = 20 µm, the data are the continuation of the data
already reported in Figure 2.C: Snapshots for each width W and two concentrations H0 = 0.6%
and 4.5%. Every snapshot is the superimposition of 10 successive images in order to highlight the
structure of the suspension in the parent vessel
4 Discussion and conclusions
As a result of the interplay between the Zweifach-Fung effect and the Fahraeus effect the hemat-
ocrit in microcirculation can reach values as low as 10-20% compared to the average hematocrit
in human body (45%). At such a low hematocrit, our simulations and in vitro microfluidic ex-
periments have revealed that RBCs partition at the level of bifurcations depends strongly on the
viscosity contrast between the viscosities of the RBC hemoglobin and the suspending fluid. In
the extreme hemodilution, our results exhibit a newly reported phenomenon: The low flow rate
branch may receive higher hematocrit than the high flow rate branch in opposition to the known
Zweifach-Fung effect. This phenomenon is observed under moderate confinement and is the result
of a peculiar structuring of the cell suspension. Under stronger confinement, other strong discrep-
ancy with Pries et al. empirical law was highlighted, with a strong asymmetry in the partitioning
12
λ=10.3λ=10.3ABC00.020.040.060.0800.020.040.06H1H0W=30µm,λ=0.85W=30µm,EmpiricalW=20µm,λ=0.85W=20µm,EmpiricalW=10µm,λ=0.85W=10µm,Empirical00.020.040.060.0800.020.040.06H1H0W=30µm,λ=10.3W=30µm,EmpiricalW=20µm,λ=10.3W=20µm,EmpiricalW=10µm,λ=10.3W=10µm,EmpiricalW=30µmW=20µmW=10µmλ=0.85λ=0.85H0=0.6%H0=4.5%H0=0.6%H0=4.5%(enhanced Zweifach-Fung effect). Our findings suggest that the viscosity contrast must be taken
into consideration and carefully analyzed in order to have a firm understanding of RBC distribu-
tion in microcirculation. This physiological parameter increases with aging as well as with some
pathologies.
The results of our present work provide a valuable background needed to pinpoint the various
RBCs properties that govern hematocrit partition, and thus oxygen delivery in the microcirculation
in general.
Acknowledgment
Z. S. thanks Vassanti Audemar for experimental advice and assistance. Z. S., G. C., C. M. and
T. P. acknowledge financial support from CNES (Centre National d’Etudes Spatiales) and ESA
(European Space Agency). The DyFCom team of LIPhy (Z. S., G. C., C. M. and T. P.) is mem-
ber of Labex TEC21 (Investissements d’Avenir-Grant Agreement ANR-11-LABX-0030), Struc-
ture F´ed´erative de Recherche iDYSCO (CNRS), F´ed´eration Galileo Galilei de Grenoble (FR3345
CNRS-UJF-Grenoble INP-IRSTEA) and Groupement de Recherche MECABIO (GDR3570 CNRS).
References
M. Abkarian, M. Faivre, and A. Viallat. Swinging of red blood cells under shear flow. Phys. Rev.
Lett., 98:188302, 2007.
D. M. Audet and W. L. Olbricht. The motion of model cells at capillary bifurcations. Microvasc.
Res., 33:377–396, 1987.
J. O. Barber, J. P. Alberding, J. M. Restrepo, and T. W. Secomb. Simulated two-dimensional red
blood cell motion, deformation, and partitioning in microvessel bifurcations. Ann. Biomech.
Eng., 36:1690–1698, 2008.
N. Callens, C. Minetti, G. Coupier, M. Mader, F. Dubois, C. Misbah, and T. Podgorski. Hydrody-
namic lift of vesicles under shear flow in microgravity. Europhys. Lett., 83:24002, 2008.
I. Cantat and C. Misbah. Lift force and dynamical unbinding of adhering vesicles under shear flow.
Phys. Rev. Lett., 83:880–883, 1999.
S. Chien, C. D. Tvetenstrand, M. A. Epstein, and G. W. Schmid-Schonbein. Model studies on
distributions of blood cells at microvascular bifurcations. Am. J. Physiol. Heart Circ. Physiol.,
248:H568–H576, 1985.
G. R. Cokelet and H. J. Meiselman. Rheological comparison of hemoglobin solutions and erythro-
cyte suspensions. Science, 162:275–377, 1968.
D. V. Cokkinos, C. Pantos, G. Heusch, and H. Taegtmeyer. Myocardial ischemia:
from
mechanisms to therapeutic potentials, volume 21. Springer Science & Business Media, 2006.
13
G. Coupier, B. Kaoui, T. Podgorski, and C. Misbah. Noninertial lateral migration of vesicles in
bounded Poiseuille flow. Phys. Fluids, 20:111702, 2008.
J. W. Dellimore, M. J. Dunlop, and P. B. Canham. Ratio of cells and plasma in blood flowing past
branches in small plastic channels. Am. J. Physiol. Heart Circ. Physiol., 244:H635–H643, 1983.
R. Ditchfield and W. L. Olbricht. Effects of particle concentration on the partitioning of suspen-
sions at small divergent bifurcations. J. Biomech. Eng., 118:287–294, 1996.
V. Doyeux, T. Podgorski, S. Peponas, M. Ismail, and G. Coupier. Spheres in the vicinity of a
bifurcation: elucidating the Zweifach-Fung effect. J. Fluid Mech., 674:359, 2011.
J. Dupire, M. Socol, and A. Viallat. Full dynamics of a red blood cell in shear flow. Proceedings
of the National Academy of Sciences, 109(51):20808–20813, 2012.
A. Farutin and C. Misbah.
Symmetry breaking and cross-streamline migration of three-
dimensional vesicles in an axial poiseuille flow. Physical Review E, 89(4):042709, 2014.
D. Fedosov, B. Caswell, S. Suresh, and G. Karniadakis. Quantifying the biophysical characteristics
of plasmodium-falciparum-parasitized red blood cells in microcirculation. Proceedings of the
National Academy of Sciences, 108(1):35–39, 2011.
B. M. Fenton, R. T. Carr, and G. R. Cokelet. Nonuniform red cell distribution in 20 to 100 µm
bifurcations. Microvasc. Res., 29:103–126, 1985.
Y.-C. Fung. Biomechanics: circulation. Springer Science & Business Media, 2013.
X. Grandchamp, G. Coupier, A. Srivastav, C. Minetti, and T. Podgorski. Lift and down-gradient
shear-induced diffusion in red blood cell suspensions. Phys. Rev. Lett., 110:108101, 2013.
R. Guibert, C. Fonta, and F. Plouraboue. A New Approach to Model Confined Suspensions Flows
in Complex Networks: Application to Blood Flow. Transport Porous Med., 83:171–194, 2010.
D. S. Hariprasad and T. W. Secomb. Two-dimensional simulation of red blood cell motion near a
wall under a lateral force. Phys. Rev. E, 90:053014, 2014.
T. Hyakutake and S. Nagai. Numerical simulation of red blood cell distributions in three-
dimensional microvascular bifurcations. Microvasc. Res., 97:115 – 123, 2015. ISSN 0026-2862.
B. Kaoui, J. Harting, and C. Misbah. Two-dimensional vesicle dynamics under shear flow: Effect
of confinement. Physical Review E, 83(6):066319, 2011.
D. Katanov, G. Gompper, and D. A. Fedosov. Microvascular blood flow resistance: Role of red
blood cell migration and dispersion. Microvascular research, 99:57–66, 2015.
C. Kelemen, S. Chien, and G. Artmann. Temperature transition of human hemoglobin at body
temperature: effects of calcium. Biophysical journal, 80(6):2622–2630, 2001.
T. Kruger, B. Kaoui, and J. Harting. Interplay of inertia and deformability on rheological properties
of a suspension of capsules. Journal of Fluid Mechanics, 751:725–745, 2014.
14
A. Kumar, R. G. Henr´ıquez Rivera, and M. D. Graham. Flow-induced segregation in confined
multicomponent suspensions: effects of particle size and rigidity. J. Fluid Mech., 738:423–462,
1 2014. ISSN 1469-7645.
X. Li, A. S. Popel, and G. E. Karniadakis. Blood-plasma separation in y-shaped bifurcating mi-
crofluidic channels: a dissipative particle dynamics simulation study. Phys. Biol., 9:026010,
2012.
J. L. McWhirter, H. Noguchi, and G. Gompper. Flow-induced clustering and alignment of vesicles
and red blood cells in microcapillaries. Proceedings of the National Academy of Sciences, 106
(15):6039–6043, 2009.
S. T. Ollila, C. Denniston, and T. Ala-Nissila. One- and two-particle dynamics in microfluidic
t-junctions. Phys. Rev. E, 87:050302, 2013.
C. S. Peskin. The immersed boundary method. Acta numerica, 11:479–517, 2002.
A. R. Pries, K. Ley, M. Claassen, and P. Gaethgens. Red cell distribution at microvascular bifur-
cations. Microvasc. Res., 38:81–101, 1989.
A. R. Pries, T. W. Secomb, P. Gaehtgens, and J. F. Gross. Blood flow in microvascular networks.
experiments and simulation. Circ. Res., 67(4):826–34, 1990. doi: 10.1161/01.RES.67.4.826.
B. W. Roberts and W. L. Olbricht. Flow-induced particulate separations. AIChE J., 49:2842–2849,
2003.
B. W. Roberts and W. L. Olbricht. The distribution of freely suspended particles at microfluidic
bifurcations. AIChE J., 52:199–206, 2006.
J. M. Sherwood, E. Kaliviotis, J. Dusting, and S. Balabani. Hematocrit, viscosity and ve-
locity distributions of aggregating and non-aggregating blood in bifurcating microchannel.
Biomicrofluidics, 6:024119, 2012.
L. Shi, T.-W. Pan, and R. Glowinski. Numerical simulation of lateral migration of red blood cells
in poiseuille flows. Int. J. Numer. Methods Fluids, 68(11):1393–1408, 2012.
K. Tsubota, S. Wada, and T. Yamaguchi. Particle method for computer simulation of red blood cell
motion in blood flow. Computer methods and programs in biomedicine, 83(2):139–146, 2006.
H. C. Woolfenden and M. G. Blyth. Motion of a two-dimensional elastic capsule in a branching
channel flow. J. Fluid Mech., 669:3–31, 2 2011.
W. Xiong and J. Zhang. Two-dimensional lattice boltzmann study of red blood cell motion through
microvascular bifurcation: cell deformability and suspending viscosity effects. Biomech. Model.
Mechanobiol., 11:575–583, 2012.
X. Yin, T. Thomas, and J.Zhang. Multiple red blood cell flows through microvascular bifurcations:
Cell free layer, cell trajectory, and hematocrit separation. Microvasc. Res., 89:47 – 56, 2013.
15
J. Zhang, P. C. Johnson, and A. S. Popel. An immersed boundary lattice boltzmann approach to
simulate deformable liquid capsules and its application to microscopic blood flows. Physical
biology, 4(4):285, 2007.
16
|
1109.5786 | 1 | 1109 | 2011-09-27T06:45:56 | Growth of bioluminescent bacteria under modelled gravity of different astronomical bodies in the Solar system | [
"physics.bio-ph",
"q-bio.CB"
] | Spaceflights and clinostats have been used extensively to study the effects of microgravity on various biological systems ranging from microbes to plants. Similarly hypergravity studies have been carried out using centrifuges where growth retardation has been observed. However, no studies have been carried out yet on how the gravity of astronomical bodies, e.g. Moon having 1/6th the gravity of Earth, affects biological systems. Such studies are important with missions to Moon and Mars to be carried out in future. Also, a comparative study to see the effects of gravity that exists on astronomical bodies such as Moon, Mars and Jupiter on any organism using simulation have not been reported so far. This paper discusses the effects of modelled gravity on the growth of Vibrio harveyi using the clinostat-centrifuge system designed and developed in-house. Results showed that though growth as measured by optical density was significantly higher for simulated microgravity and lunar and Martian gravities, there was no significant difference in viable counts. This is because the relative death rate is also higher for these gravities. Jovian gravity was found to slightly retard the growth. This study also shows that simulated lunar gravity is relatively most suited for the growth of Vibrio harveyi. | physics.bio-ph | physics | Growth of bioluminescent bacteria under modelled gravity of
different astronomical bodies in the Solar system
Santosh Bhaskaran*1,@, Rohan Dudhale1,2,$, Jyotsana Dixit1,2, Ajit Sahasrabuddhe1,3,# and
Pandit B. Vidyasagar1
1Biophysics Laboratory, Department of Physics, University of Pune, Pune 411 007, INDIA
2Department of Microbiology, P.E.S. Modern College of Arts, Science and Commerce, Ganeshkhind, Pune
411 053 INDIA. $ Haffkine Institute for Training, Research and Testing, Parel, Mumbai 400 012, INDIA
3Department of Microbiology, P.E.S. Modern College of Arts, Science and Commerce, Shivajinagar, Pune
411 005, INDIA #Quality Control Division, Serum Institute of India, Hadapsar, Pune 411 028, INDIA
Abstract
Spaceflights and clinostats have been used extensively to study the effects of
microgravity on various biological systems ranging from microbes to plants. Similarly
hypergravity studies have been carried out using centrifuges where growth
retardation has been observed. However, no studies have been carried out yet on how
the gravity of astronomical bodies, e.g. Moon having 1/6th the gravity of Earth, affects
biological systems. Such studies are important with missions to Moon and Mars to be
*Biophysics Laboratory, Department of Physics, University of Pune, Pune 411 007, INDIA
Fax: +91-44-22232711
Email: [email protected]
Present Address
@Life Sciences Division, AU-KBC Research Centre, MIT Campus, Chennai 600 044, INDIA
carried out in future. Also, a comparative study to see the effects of gravity that exists
on astronomical bodies such as Moon, Mars and Jupiter on any organism using
simulation have not been reported so far. This paper discusses the effects of modelled
gravity on the growth of Vibrio harveyi using the clinostat-centrifuge system designed
and developed in-house. Results showed that though growth as measured by optical
density was significantly higher for simulated microgravity and lunar and Martian
gravities, there was no significant difference in viable counts. This is because the
relative death rate is also higher for these gravities. Jovian gravity was found to
slightly retard the growth. This study also shows that simulated lunar gravity is
relatively most suited for the growth of Vibrio harveyi.
Keywords: Microgravity, Hypogravity, Hypergravity, Clinostat-centrifuge, Moon, Jupiter
Running Title: Growth of Vibrio harveyi in altered gravity
2
Introduction
Experiments on Escherichia coli in space showed a shortened lag phase, an increased duration of
exponential growth and an approximate doubling of final cell population density compared to controls (Klaus
et al., 1997; Gasset et al., 1994) but no change in growth rate was observed (Brown et al., 2002). Modelled
microgravity also showed an enhanced growth of E. Coli while hypergravity retarded the growth (Gasset et
al., 1994). Bacillus subtilis showed a higher growth rate in microgravity (Mennigmann & Lange, 1986;
Kacena et al., 1999) as well as higher final biomass yield (Mennigmann & Lange, 1986). Bacillus subtilis also
showed a shortened lag phase during spaceflight at 23 °C but not at 37 °C (Kacena et al., 1999). Salmonella
typhimurium showed a higher growth rate and reduction in generation time under modelled microgravity
(Wilson et al., 2002). An increased growth was observed in spaceflight as well (Mattoni, 1968). Modelled
microgravity decreased the germination efficiency of Dictyostelium discoideum while hypergravity promoted
it (Kawasaki et al., 1990). Fruiting bodies were found to be smaller in modelled microgravity and taller in
hypergravity (Kawasaki et al., 1990). Paramecium showed a random swimming behaviour below 0.16g with
negative gravitaxis becoming pronounced from 0.3 g suggesting a threshold for gravitaxis between 0.16 and
0.3 g (Hemmersbach et al, 1996). Studies on the contraction activity of Physarum onboard SPACELAB-I
showed that the threshold was 0.1 g to elicit a response (Block et al, 1996).
In this paper, we discuss how modelled gravity conditions of Moon (0.16g), Mars (0.38g) and Jupiter
(2.5g) affects the growth of Vibrio harveyi, a bioluminescent bacterium, isolated from the coastal waters of
Goa, India. Bioluminescent bacteria have applications as environmental biosensors and pollution indicators
(Girotti et al., 2008). We show that though microgravity, lunar and Martian gravities enhance the growth, they
also enhance the death of Vibrio harveyi. To the best of our knowledge, neither studies on the effects of lunar
and Martian gravities (hypogravity) nor a comparative study of different gravity environments have been
carried out on the growth of any organism.
3
Materials and Methods
Bacterial cultures obtained from coastal waters of Goa, India were isolated and identified as Vibrio
harveyi (strain NB0903) using 16S rRNA
sequencing. The
suspension culture was
adjusted to OD = 0.2 (approx. 5x109 cells/ml)
and 1% inoculum added to sterile BOSS broth
(Klein et al., 1998) in each of four identical
glass vessels (8 cm x 11 cm), specially designed
for the clinostat-centrifuge system (designed
and developed in-house) (Fig.1). Two of these
cultures were exposed to modelled varied
gravity conditions viz. Microgravity (Space),
0.16 g (Lunar), 0.38 g (Martian), 1 g (Earth) as
dynamic control and 2.5 g (Jovian) for 15 hours. The g-values were calculated using the standard formula
RCF = 1.118x10-5 x r x N2….
(1)
where
RCF is the relative centrifugal force in g units,
R (=17 cm) is the distance of the sample from the centre of the clinostat-centrifuge in cm. This includes the
distance of the sample holder from the centre and half the length of the vessel as the rotation is along the
vertical axis. In case of microgravity r (= 4 cm) is the radius of the vessel as the rotation is along the
horizontal axis.
and
N is the speed of rotation in rpm
4
In a normal centrifuge, two forces viz., gravitational force downward and the centrifugal force outwards, act
on the sample. The net force acting on the sample is the resultant of the two and hence the magnitude of the
resultant would be always higher than 1g. However in a clinostat, the sample is in modelled microgravity.
Thus in a clinostat-centrifuge system, accelerations less than 1g can also be modelled.
The other two cultures acted as static controls. Static controls were used since shaking at 170 rpm
would be equivalent to 3g. From the time of inoculation, at regular intervals of one hour, aliquots of 2 ml
were taken for measuring optical density at 590 nm using a digital colorimeter and for total viable count using
spread plate method in duplicates. Both control and test vessels were shaken well before taking the aliquots so
that the cultures were well-mixed. Changes, if any, in morphological characteristics were also observed.
Since OD is a function of the total cell number for a particular instant, the difference in OD and TVC
can be correlated to the number of dead cells at that instant. The relative death rates for different values of g
were calculated as follows. Log CFU versus OD in the exponential phase was plotted and a mathematical
function relating them was obtained. Taking this relation as a standard, the total number of cells in the broth
for each value of g was estimated by substituting the corresponding OD values in this relation. The difference
between the total number of cells in broth and CFU gives us the relative number of dead cells. This was used
to obtain the relative death rate. Since the relation for static control is taken as zero, the relative death rate
would be zero.
All experiments were carried out at room temperature (25±1 °C). Each experiment was repeated
three times and consistent results were obtained. All data are represented as Mean ± SEM. The p-values were
obtained by Student’s T-test.
5
6
plates have shown similar results (Kacena et al., 1997). Similar results were obtained for Pseudomonas
aeruginosa grown under modelled microgravity (Guadarrama et al., 2005). Some studies have attributed this
to the influence of motility (Thevenet et al., 1996, Benoit & Klaus, 2007) while some others have attributed
this to lack of more efficient metabolic capabilities in microgravity (Kacena et al., 1997). It should however
be noted that in our case, liquid cultures were grown under altered gravity while viable counts for all the
treated cultures were obtained under normal gravity on agar plates using the aliquots taken from these liquid
cultures. Yet similar results have been obtained.
From Table 1, it can be seen that growth rate from OD was the highest for modelled lunar gravity.
This was followed by the growth rate for modelled microgravity and modelled Martian gravity which were
almost the same while Jovian gravity showed the minimum growth rate. However, the relative death rate is
also higher in modelled microgravity and lunar and Martian gravities. This shows why the viable cell counts
are not significantly different from control. Interestingly, the relative death rate is lesser than the growth rate
from OD only for lunar gravity. This is reflected in the form of least generation time and maximum growth
rate from viable cell counts. This shows that in effect; modelled microgravity, lunar and Martian gravities
enhance the growth of Vibrio harveyi with modelled lunar gravity showing maximum enhancement while
Jovian gravity retards the growth. The lag phase was reduced by 50 % for modelled microgravity, lunar and
Martian gravities but remained the same for Jovian gravity with respect to control. As far as the exponential
phase is concerned, it increased by 50 % in modelled microgravity, 33 % in modelled lunar gravity and 17 %
in modelled Martian gravity but remained the same in Jovian gravity with respect to control.
No changes in colony morphological characteristics were observed for all samples exposed to each of
the altered gravity conditions. SDS-PAGE did not show any change in band patterns for all samples (data not
shown).
We expected an inverse relation between the growth and gravity and indeed it is the case except for
lunar gravity. Hence it was quite surprising to get maximum growth for lunar gravity instead of microgravity.
These results can be concluded due to gravity alone and not due to other factors such as aeration since growth
for Jovian is less than 1g control. If it was the case, then maximum growth should have been observed for
7
Jovian gravity since the rotation speed is maximum and hence the aeration relative to all other values of g
reported in this paper.
In summary, we have measured the growth rates by OD and viable counts under different modelled
gravity conditions, however a significant difference is found only in the OD. Organism under modelled
microgravity, lunar and Martian gravity conditions shows, higher growth rate and reduced generation time
with respect to static controls while Jovian gravity did not have a significant impact on the organism studied.
Non-significant difference in viable counts is due to an equal number of cells dying under these conditions
and not possibly due to similar metabolic activities (Kacena et al., 1997) in microgravity or differences in
motility (Benoit & Klaus, 2007).
We have shown that modelled lunar gravity, i.e. 0.16g, is most favourable for the growth of Vibrio
harveyi. However, it is important to know if our results are a general trend for all organisms or are typical of
Vibrio harveyi alone. Similar studies on mammalian cells are important, since they will provide an insight as
to how our physiological systems will adapt to such gravity conditions. This has great implications for setting
up bases on other planets and their moons. This can also be used in industry to get maximum output from
micro-organisms just by optimising the value of g in addition to pH, temperature, etc.
References
Benoit, M.R., and Klaus, D.M. (2007) Microgravity, bacteria, and the influence of motility. Adv. Space Res.
39, 1225-1232.
Block I., Briegleb W. and Wolke A. (1996) Acceleration-sensitivity threshold of Physarum. J. Biotechnol. 47,
239-244
Brown, R.B., Klaus, D. And Todd, P. (2002) Effects of space flight, clinorotation, and centrifugation on the
substrate utilization efficiency of E. Coli. Micrograv. Sci. Tech. 13, 24-29.
8
Gasset, G., Tixador, R., Eche, B., Lapchine, L., Moatti, N., Toorop, P., and Woldringh, C. (1994) Growth and
division of Escherichia coli under microgravity conditions. Res. Microbiol. 145, 111-120.
Girotti, S., Ferri, E.N., Fumo, M.G., and Maiolini, E. (2008) Monitoring of environmental pollutants by
bioluminescent bacteria. Anal. Chim. Acta. 608, 2-29
Guadarrama, S., Pulcini, E.L., Broadaway, S.C. and Pyle, B.H. (2005) Pseudomonas aeruginosa growth and
production of Exotoxin A in static and modeled microgravity environments. Gravit. Space Biol. 18, 85-86.
Hemmersbach R., Voormanns R.,, Briegleb W., Riederb N. and Hider D.P. (1996) Influence of accelerations
on the spatial orientation of Loxodes and Paramecium. J. Biotechnol. 47, 271-278
Kacena, M.A., Leonard, P.E. Todd, P, and Luttges, M.W. (1997) Low gravity and inertial effects on the
growth of E. Coli and B. Subtilis in semi-solid media. Aviat. Space Environ. Med. 68, 1104-1108.
Kacena, M.A., Merrell G.A., Manfredi B., Smith E.E., Klaus D.M. and Todd P. (1999) Bacterial growth in
spaceflight: Logistic growth curve parameters for Escherichia coli and Bacillus subtilis. Appl.
Microbiol.Biotechnol. 51, 229-234.
Kawasaki, Y., Kiryu, T., Usui, K. And Mizutani, H. (1990) Growth of the cellular slime mold, Dictyostelium
discoideum, Is gravity dependent. Plant Physiol. 93, 1568-1572.
Klaus, D., Simske, S., Todd, P. And Stodieck, L. (1997) Investigation of space flight effects on Escherichia
coli and a proposed model of underlying physical mechanisms. Microbiol. 143, 449-455.
Klein, G., mijewski, M., Krzewska, J., Czeczatka, M. And Lipi ska, B. (1998) Cloning and characterization
of the dnak heat shock operon of the marine bacterium Vibrio harveyi. Mol. Gen. Genet. 259, 179-189.
Mattoni, R.H.T. (1968) Space flight effects and gamma radiation interaction on growth and induction of
lysogenic bacteria. Bioscience. 18, 602-608.
9
Mennigmann, H.D. and Lange, M. (1986) Growth and differentiation of Bacillus subtilis under microgravity.
Naturwissenschaften 73, 415417.
Thevenet, D., D’Ari, R. And Bouloc, P. (1996) The SIGNAL experiment in BIORACK: Escherichia coli in
microgravity. J. Biotech. 47, 89-97.
Wilson, J.W., Ott, C.M., Ramamurthy, R., Porwollik, S., mcclelland, M., Pierson, D.L., and Nickerson, C.A.
(2002) Low-Shear modeled microgravity alters the Salmonella enterica serovar typhimurium stress response
in an rpos-independent manner. Appl. Environ. Microbiol. 68, 5408– 5416.
Acknowledgements
We are thankful to the Director, National Centre for Cell Science, Pune, India as well as to Prof. B.A.
Chopade and Mr. .Praveen K. Sahu, DNA sequencing Laboratory, Institute of Bioinformatics &
Biotechnology, University of Pune, India for carrying out the 16S rrna sequencing of the bacterial isolate.
Comments provided by Mr. Vivek Jadhav, University of Pune, Mr. Anant Rajeha, I.I.T. Chennai, Mr. Lasse
Folkerson, Karolinska Institute, Stockholm and Ms. Jamila Siamwala, AU-KBC Research Centre, Chennai
were of great help in writing this paper. Co-author PBV would like to thank abdus-Salam International Centre
of Theoretical Physics, Trieste for Associateship.
Author Contributions SB, AS and RD conceived the idea of this work. AS and SB made designs of the
vessels while RD and JD isolated and characterised the organism. SB and AS designed the experiments. RD,
JD, AS and SB contributed equally to the growth curve experiments. JD carried out the SDS PAGE. SB and
JD mainly wrote the paper. PBV supervised this work.
10
Table 1. Growth kinetics in altered gravity
g
Growth Rate
Relative
Growth Rate
Generation
Lag
Exponential
from OD
Death Rate
from TVC
Time
Phase
Phase
(Hour-1)
(Hour-1)
(Hour-1)
(Minutes)
(Hours)
(Hours)
Static Control (1g)
0.365 ± 0.038
-
0.365 ± 0.038
114.9 ± 11.12
Space (µg)
0.51 ± 0.043
0.651 ± 0.174
0.392 ± 0.043 107.01 ± 11.75
Moon (0.16g)
0.602 ± 0.044
0.57 ± 0.232
0.41 ± 0.056
102.65 ± 12.56
Mars (0.38g)
0.432 ± 0.042
0.485 ± 0.171
0.394 ± 0.037
106.38 ± 9.65
Dynamic Control (1g)
0.294 ± 0.042
0.436 ± 0.159
0.383 ± 0.105 114.92 ± 12.83
Jupiter (2.5g)
0.235 ± 0.041
0.258 ± 0.128
0.34 ± 0.037
123.26 ± 12.89
2
1
1
1
2
2
6
9
8
7
6
6
11
|
1704.05885 | 1 | 1704 | 2017-04-19T18:27:10 | Integrating optimization with thermodynamics and plant physiology for crop ideotype design | [
"physics.bio-ph"
] | A computational framework integrating optimization algorithms, parallel computing and plant physiology was developed to explore crop ideotype design. The backbone of the framework is a plant physiology model that accurately tracks water use (i.e. a plant hydraulic model) coupled with mass transport (CO2 exchange and transport), energy conversion (leaf temperature due to radiation, convection and mass transfer) and photosynthetic biochemistry of an adult maize plant. For a given trait configuration, soil parameters and hourly weather data, the model computes water use and photosynthetic output over the life of an adult maize plant. We coupled this validated model with a parallel, meta-heuristic optimization algorithm, specifically a genetic algorithm (GA), to identify trait sets (ideotypes) that resulted in desired water use behavior of the adult maize plant. We detail features of the model as well as the implementation details of the coupling with the optimization framework and deployment on high performance computing platforms. We illustrate a representative result of this framework by identifying maize ideotypes with optimized photosynthetic yields using weather and soil conditions corresponding to Davis, CA. Finally, we show how the framework can be used to identify broad ideotype trends that can inform breeding efforts. The developed presented tool has the potential to inform the development of future climate-resilient crops. | physics.bio-ph | physics | Integrating optimization with thermodynamics and plant physiology for crop
ideotype design
Talukder Z. Jubery1, Baskar Ganapathysubramanian1, Matthew E. Gilbert2, and Daniel
Attinger1
1Department of Mechanical Engineering and Department of Electrical and Computer
Engineering, Iowa State University, Ames, IA, 50011;
2Department of Plant Sciences, University of California, Davis, CA 95616, USA.
*Correspondence to: [email protected], [email protected], [email protected]
Abstract
A computational framework integrating optimization algorithms, parallel computing and plant
physiology was developed to explore crop ideotype design. The backbone of the framework is a
plant physiology model that accurately tracks water use (i.e. a plant hydraulic model) coupled
with mass transport (CO2 exchange and transport), energy conversion (leaf temperature due to
radiation, convection and mass transfer) and photosynthetic biochemistry of an adult maize plant.
For a given trait configuration, soil parameters and hourly weather data, the model computes
water use and photosynthetic output over the life of an adult maize plant. We coupled this
validated model with a parallel, meta-heuristic optimization algorithm, specifically a genetic
algorithm (GA), to identify trait sets (ideotypes) that resulted in desired water use behavior of the
adult maize plant. We detail features of the model as well as the implementation details of the
coupling with the optimization framework and deployment on high performance computing
platforms. We illustrate a representative result of this framework by identifying maize ideotypes
with optimized photosynthetic yields using weather and soil conditions corresponding to Davis,
CA. Finally, we show how the framework can be used to identify broad ideotype trends that can
inform breeding efforts. The developed presented tool has the potential to inform the
development of future climate-resilient crops.
Keywords: ideotypes, optimization, net photosynthesis, hydraulic traits.
Introduction
To ensure food security, crop grain yields should be increased globally by 70–100% within the
next 40 years [1]. To increase yields, plant breeders and plant scientists are working to develop
improved and appropriate varieties of crops. However, the intrinsic uncertainty of climate
change, limited water supply and reduction of agricultural land increase the challenges in the
crop development process [2]. We have limited time and resources to select the most appropriate
crop varieties, and crop modeling provides a rational approach to designing new crop varieties
[3].
Traditional methods for finding the best crop varieties, or ideotypes [4], rely on agronomic
experiments. The evaluated ideotypes are restricted in time and space, making results site- and
1
season-specific, and the experiments are time consuming and expensive. The use of crop models
has greatly enabled crop breeding by reducing costs and accelerating the process of
identifying/designing ideotypes.
Many physiological models of crops [5]–[15] have been developed since the pioneering 1948
model of van den Honert [16]. These models are focused on specific aspects of plant physiology:
water transport, time-dependence, influence of environmental conditions, heat and mass
transfer, effect of plant geometry, nutrient transport, plant growth, or phloem transport. While
these models can determine and explain optimum relationships between existing traits [5], there
is increasing interest in coupling them with optimization tools to identify the most promising
traits for a desired response [24], [25]. Over the last three decades, publications on numerical
optimization methods have emerged and their number has grown at a rate higher than the growth
rate of publications on traditional plant breeding. Considering also that breeding is, per se, an
optimization problem, the recent emergence of limited publications at the intersection of
"breeding" and "numerical optimization" is not unexpected. Some of these publications describe
numerical methods inspired by animal breeding strategies [17], [18], while others seek to
optimize the management of a breeding program [19], [20] , either by improving the phenotyping
associated with breeding [21], or minimizing the genotyping efforts [22]. While these models
can determine and explain optimum relationships between existing traits [5], [10], [23], there is
work yet to be done to efficiently leverage this knowledge to direct breeding efforts. This is the
motivation of our work. We describe our model and how we integrate it with an optimization
framework. We then demonstrate an application of the framework by identifying crop ideotypes
for a specific location parameterized by weather and soil conditions.
Materials and Methods
Plant physiology model
The backbone of our framework is a mechanistic crop physiological model that is based upon a
detailed one-dimensional representation of plant hydraulic characteristics. Liquid-phase, plant-
water relations are simply represented as a static series of conductances resistances for stems,
leaves and roots ( disregarding capacitive behavior, i.e. stems, leaves and roots do not store any
water) , as described in the seminal work of van den Honert [16]. The model is a physiologically
explicit representation of C4 maize water-use after canopy closure. The model explicitly
accounts for energy balance (convection, radiation, latent heat), transpiration, intercellular CO2
concentration (via both diffusion and biochemical processes), weather conditions (temperature,
precipitation, pressure, radiation), and soil type. Seven plant hydraulic 'traits' are considered
within the model, as shown in Fig. 1, and can be used to represent the response of leaf
evapotranspiration to environmental variation. We next provide details of each submodel that is
used to construct the full plant model. Figure 2 shows the various submodels schematically.
2
Water transport submodel
Uptake of water by the root from the soil reaches the top of the canopy due to cohesive-adhesion
interactions, but the main driving force for this transport is the dryness of the atmosphere. Water
travels from the soil to the leaf as a liquid. Subsequently, as a gas it evaporates from the leaf
(through the stomatal pores) to the surrounding environment. The evaporation rate depends on
the leaf temperature, external relative humidity, air temperature and boundary layer effects. This
is called environmental water demand. To fulfill this demand, the plant supplies liquid water to
the leaf. This flow is driven by the potential difference of water between the soil and the leaf and
is controlled by the hydraulic resistance of the plant.
Using a one-dimensional representation of plant hydraulic characteristics, as shown in Fig. 1,
the water supply,
, can be expressed as [26]
,
(1)
where Kplant is hydraulic conductance of the plant, Ksoil is hydraulic conductance of soil; ψleaf is
the water potential at the leaf, ψsoil is water potential at the soil. Water potential is a combined
effect of hydrostatic pressure, osmotic pressure, matric pressure, and gravitational pull.1
The hydraulic conductance of the plant can be expressed as
,
(2)
where Kroot is the hydraulic conductance of the root, Kstem is hydraulic conductance of the stem,
and Kleaf is hydraulic conductance of the leaf.
The hydraulic conductance of the soil, Ksoil, depends on the type of soil, the amount of water in
the soil and the relative occupancy of the root in the soil. The effect of these parameters is
captured via the following equation [27],
,
(3)
1 . Osmotic pressure that depends on the presence of ions in the water is neglected in our model,
as we considered the water as pure and free from any minerals.
3
sWJ,leafsoilsoilplantsWKKJ1,111111leafstemrootplantKKKKrootsbrootsoilsoilsatsatsoilrLLHkKlog2/32,where ksat, b, ψsat and ψsoil vary among the types of soil, and they represent saturated hydraulic
conductivity, texture, water potential of saturated soil, and water potential of the soil,
respectively. The rest of the terms are used to capture the effect of the presence of root on the
soil conductance. The symbols L, Hs and rroot represent root length density of the absorbing root
(length per soil volume), depth of the soil occupied by the root and radius of the root.
The water potential of the soil, ψsoil, can be expressed as a function of soil water content using an
empirical equation developed by Campbell et al. [28] as
,
(4)
where θsat is the saturated water content in the soil, and θ is current volumetric water content in
the soil. In this model, soil water content would gradually deplete as plants fulfill the
atmospheric water demand. The depletion of water due to evaporation of water from the soil was
not considered here. There is no addition of water in the case of drought conditions. However,
under irrigated conditions, based on the irrigation frequency, water is added to the soil until
water content reaches θsat of the soil. e.g., for irrigation frequency 7, the soil is fully saturated
every 7*24 hours.
Soil water potential at the root (Eq. 3), ψsoil,root, can be evaluated from leaf water potential and
water demand by the plants,
,
(5)
Water demand is driven by the gradient of water vapor concentration between the leaves and the
surrounding environment and is controlled by the stomatal conductance and air boundary layer
conductance. It can be expressed as [26]
,
(6)
where Pvl and Pva represent water vapor pressure in the leaf and atmosphere, respectively. Water
vapor pressures are evaluated using Tetens formula [29] ,
, where RH is the
relative humidity, c0 = 0.617 kPa, c1 =17.38, and c2= 239°C. Generally, the leaf inter-cellular
space is close to equilibrium with the cells having a relative humidity of greater than 99%, and
thus for each of calculation of evaporation we consider the leaf to be fully saturated. gst and gblc
are the stomatal conductance and boundary layer conductance to the water vapor transport,
respectively.
4
bsatsatsoilsoilplantdWleafrootsoilKKJ11,,avavlblcstdWPPPggJ1,11210,cTTciiviiecRHPBoundary layer conductance to water vapor, gblc depends on the atmospheric wind speed and the
morphology as well as the orientation of the leaf. Wind speed and leaf dimension are designated
as Uc, and d as in [28]. Conductance of water vapor through the air boundary layer on the leaf
can be considered as forced convection and can be expressed via an empirical equation. Note
that, here, contribution from the free convection is neglected, as the ratio of dimensionless
parameters Re2/Gr which reflects the forced convection/free convection is usually much greater
thanone. The empirical correlation among the dimensionless Reynolds number, Re, and Schmidt
number, Sc, and the conductance can be calculated as,
,
(7)
where
;
; α=0.644*1.4 is an empirical parameter; and de =0.72d, with d
being the width of the maize leaf and 0.72 being used to find the equivalent parabola of the leaf
where wind is flowing in the width direction of the parabola. Uc , νa and Dwv represent the wind
speed on the top of the canopy, kinetic viscosity of air and water vapor diffusivity in air.
Wind speed can increase approximately logarithmically with distance above a plant canopy, and
is also influenced by the plants. The variation in wind speed can be described by
,
(8)
where 0.4 is related to the von Karman constant, Hc is the height of the plant, mHc is the zero-
plane displacement, and nHc is the roughness length. Generally, m is 0.7 and n is 0.1. U* is
termed the shearing or friction velocity and can be calculated from the wind speed Um that is
measured at height Hm from the ground as
.
(9)
Only around 3% of water that is absorbed from the soil is used by the plant for
metabolism/growth, and less than 0.1% is used for photosynthesis.
CO2 transport and net photosynthesis submodel
Along with water, the plant needs CO2, sunlight and enzymes for photosynthesis. From the
environment, gaseous CO2 diffuses into the leaf via stomata and then dissolves in water and
diffuses to the cells where photosynthesis takes place. The consumption of CO2 during
photosynthesis depends on the sunlight and enzyme activity (plants always have sufficient water
to split in photosynthesis).
5
eWVblcdScDg3/12/1ReaecdUReWVaDScccccnHmHHUUln4.0*ccmmnHmHHUUln4.0*The rate of gaseous CO2 transport to the leaf is named as CO2 supply. The supply is driven by the
CO2 concentration gradient between the atmosphere and the leaf inter-cellular space, and is
controlled by the conductance of stomata and the air boundary layer. This supply can be
expressed as [26]
,
(10)
where β and χ are the ratio of CO2 conductance and water vapor conductance through stomata
and air boundary layer, respectively. β is the ratio of the molecular diffusivities of H2O and CO2,
χ is power ¾ of β, and CC,a and CC,i are the concentration of CO2 at the atmosphere and inside the
intercellular space of the leaf.
The demand of atmospheric CO2 depends on the supply of sunlight and the performance of the
enzymes that control photosynthetic activity. The plant gets some CO2 as a byproduct of
metabolism or respiration activity in the mitochondria and lowers the atmospheric CO2 demand.
For C4 plants, the electron transport to support CO2 reduction occurs in mesophyll (C4 cycle)
and bundle-sheath (C3 cycle) cells. If the supply of sunlight is lowered compared with enzyme
performance, which mainly occurs during
the morning, sunset, or cloudy days, the
photosynthetic rate can be expressed as [30]
,
(11)
where Je,t is the total electron transport rate is at leaf temperature, Rt is the rate of CO2 production
from respiration in the mesophyll and bundle sheath cell, and
is a fraction of total electrons
that are used by the mesophyll.
PEPCase, phosphoenolpyruvate carboxylase, and Rubisco are two enzymes that significantly
control the photosynthesis activity in C4 plants. PEP (three-carbon backbone) controls the
activity of the mesophyll cell (it catalyzes the primary carboxylation in a tissue that is close
to the external atmosphere) and Rubisco controls activity in bundle sheath cell. In the case of
no limitations on the supply of reductant to photosynthesis (higher light intensities), the
photosynthetic demand can be expressed as, [30]
,
(12)
Where the top expression in the right-hand side depends on the performance of PEPcase in the
mesophyll cell, and the bottom expression depends on the Rubisco performance in the bundle
6
iCaCblcstsCCCggJ,,1,ttedCRJxlightJ3)1()(,,xtROmmCbsCPEPdCRVRCgVenzymeJmax,,,min)(sheath cell. gC,bs is the bundle-sheath conductance to CO2 , CC,m is the concentration of CO2 in the
, CO2 concentration in inter-cellar space),
mesophyll cell (note that we assume that
=
Rm is mitochondrial respiration in the mesophyll at leaf temperature (i.e. CO2 supply from the
respiration of the mesophyll cell), Rt is the total mitochondrial respiration in the mesophyll and
bundle sheath at leaf temperature, and VRO max is the maximum rubisco carboxylation rate.
VPEP is the effective PEP carboxylation at leaf temperature. It depends on the availability of CO2
and the regeneration of PEP and can be expressed as [30]
,
(13)
where the top expression in right-hand side is related to the carboxylation rate of PEP, expressed
with the Michaelis-Menten Equation.
is the CO2 partial pressure in Mesophyll,
is the
maximum PEP carboxylation rate at leaf temperature, and Kp is the Michaelis-Menten constant
for PEP carboxylase for CO2 at leaf temperature. Note that the Michaelis-Menten constant, Kp,
refers to the concentration of CO2 at which the reaction rate is half of VPEPmax. The carboxylation
rate can be decreased if there is not enough PEP, and that depends on the VPEP,R, the PEP
regeneration rate at leaf temperature.
The temperature-dependent properties in the equations are evaluated using the following
equations [31]
,
,
,
,
,
,
,
7
(14)
(15)
(16)
(17)
(18)
(19)
(20)
mCC,iCC,RPEPpmCPEPmCPEPVKCVCV,,,maxminmCC,maxPEPVCTATBATVROROleafleafleafROeeQVV1110/)25(,10max25max,max10/)25(,10max25max,maxleafPEPTVPEPPEPQVV10/)25(,10,,,25leafRPEPTVRPEPRPEPQVV10/)25(,1025,leafPEPTVPEPPEPQVV10/)25(,1025,leafmTRmmQRR10/)25(,1025,leaftTRttQRR)15.273(00831.025,,leafbaTJJeteeJJwhere, A, B, C , Ja and Jb are physiological parameters related to the carboxylation rate and
electron transport rate. The subscript 25 in the symbols indicates the parameters at 25◦C. Hourly
Je,25 can be expressed as [30]
,
(21)
where λ is the empirical curvature factor and
. fPAR_PSII is the fraction of PAR
that contributes to the Photosystem II.
Using the photosynthesis rate of the above two limiting cases, the CO2 demand can be expressed
as [30]
,
(22)
Energy balance on leaf submodel
In the above equations many of the parameters related to leaves, for instance, water vapor
pressure, enzyme activities, etc., depend on the leaf temperature. Leaf temperature can be
evaluated by using first principles in so-called "big leaf models" [28]. Several assumptions are
considered in this model: the leaf is flat and perpendicular to the incident sunlight; leaf does not
store any energy; and energy storage ; and there is negligible heat generation due to metabolic
activity in the leaf.. Considering the leaf is at steady state, the energy balance equation on a leaf
can be expressed as [28]
,
(23)
where the terms are energy input by solar irradiation and the surrounding irradiation, cooling by
leaf irradiation, convective/conductive cooling by the air/temperature gradient and heat loss
accompanying water evaporation. In Equation (23), a is the absorptance of the leaf, r is the
reflectance, S is the solar irradiation, aIR is the absorptance of leaf for thermal infrared radiation,
Lvap is the latent heat of vaporization of water, hc is the convective heat transfer coefficient, and
ghbc is the air boundary conductance to heat transfer.
The boundary layer conductance depends on leaf morphology and wind speed, and can be
expressed via empirical relationships of dimensionless parameters Reynolds number, Re, and
Prandtl number, Pr. It can be expressed as
,
(24)
8
2425,225,25,25,maxmaxmaxeeeeIJJIJIJPSIIPARfPARI_)(),(min,,,enzymeJlightJJdCdCdC0]2[]1[,44vapdWaleafhbcpleafIRaIRLJTTgCTeTaSraeHhbcdDg3/12/1PrRewhere
;
; β=0.644*1.4 is an empirical parameter; de =0.72d, with d being the
width of maize leaf and 0.72 being used to find the equivalent parabola of the leaf where the
wind is flowing in the width direction of the parabola. Uc , νa and DH represent the wind speed on
the top of the canopy, kinetic viscosity of air and thermal diffusivity in air. The effect of the
temporal variation of soil is not explicitly included in Equation (23). Instead, the effect was
implemented using the FAO56 algorithm, as in [32].
Stomatal conductance submodel
In the pathway of the supply of CO2 (Eq. (10)) from the environment and demand of H2O (Eq.
(6)) to the environment, stomatal conductance is the most significant parameter. In general,
stomatal conductance is around several orders of magnitude lower than that for air boundary
layer conductance. Stomatal conductance is a very complex parameter that is affected by
environment, plant physiology and heredity.
At least 35 empirical models have been proposed to capture the complex relationship between
stomata conductance and various factors including [5], [8], [33]–[38] Such factors include
environmental factors, for example, solar radiation, soil water content, humidity and wind speed,
etc., and physiological factors, for example, leaf water potential, root water potential, hydraulic
root conductance, etc. Few models explicitly include the plant physiological influences on the
stomatal conductance apart from entirely empirical functions. Here, we propose a model which is
developed based on the sigmodal response of the stomatal conductance with respect to the leaf
water potential [39]. The main concept of this model is shown in Fig. 1. Here, the stomatal
conductance will start decreasing when leaf water potential touches the threshold potential,
which depends on the plant genotype. Closing rate is controlled by the two sensitivity terms Sl
and Sr and also the root water potential. The model is expressed as,
,
(25)
where the environmental response on the stomatal conductance is implicitly influenced by JC,d
and ψleaf . ψth is the threshold bulk leaf water potential at stomatal closure, and Sr is the slope of
are plant
the relationship between stomatal conductance and root water potential, ψroot.
and
physiological properties related to photosynthesis. Z is a parameter to make the exponent
dimensionless.
Method to evaluate net photosynthesis and water usage
Figure 2 shows the schematic of the concept and Fig. 3 shows the flow chart of the model
implementation. For the input weather condition, soil and agronomic/management practices the
net photosynthesis and water transpiration (Tr) can be evaluated iteratively by satisfying (Eq. 6)
9
aecdUReHaDPrZSSthleafgdCstrrootlJgggg0526.01,min2,1maxmin1g2g(Eq. 10), (Eq. 23) and (Eq. 25). A plant is considered dead and net photosynthesis is zero if the
plant experiences a permanent wilting condition or permanent temperature damage. Both states
cause irreversible damage to the plant.
Framework for crop design
Ideotype design requires identifying the optimal combination of plant physiological traits to
maximize photosynthesis for specific environmental conditions and management practices. We
formulate the design problem as an optimization problem. Thus, by writing photosynthesis as the
following functional form,
The optimization problem is defined as
,
(26)
where i represents different conditions related with agronomic practices (e.g. no-irrigation,
weekly irrigation, etc.), weather conditions, or soil type.
is a weighing factor that depends on
the preference of the designer.
Physiological traits and location/weather/management conditions
Physiological Traits: In our model, a plant has been represented by 37 physiological traits.
Typical values of most of those traits were collected from the current literature (Table 1). Note
that these traits represent the adult crop. The traits used in the photosynthetic submodel were
collected via gas exchange calibration.
Among the 37 traits in this study, we considered seven hydraulic traits: minimum stomatal
opening (
); maximum stomatal opening (
); sensitivity of stomatal opening with leaf
water potential (
); threshold bulk leaf water potential at stomatal closure (
); sensitivity of
with root water potential (
); shoot hydraulic conductance (
); and root hydraulic
conductance (
). These traits affect the stomatal conductance which is a vital trait for
photosynthesis [ref]. For the optimization problem, those traits were bounded within the ranges
in Table 2, ranges currently found in nature.
Traits related to the photosynthetic submodel: Parameters used in the photosynthetic submodel
are difficult to find in literature. Seventeen physiological parameters used in the model (that are
related to the photosynthesis equations 14-20) were calibrated using gas exchange data. The net
photosynthetic rate (An) was calibrated using gas exchange measurements made on leaves of two
maize plants grown in mini-lysimeters at the Davis Agricultural experimental station in June to
July 2013. Net photosynthetic rate was modelled as a function of three inputs: intercellular CO2
concentration, photosynthetically active radiation (PAR) and leaf temperature. Thus, these three
variables were varied using a LI-COR 6400 gas exchange system to obtain sufficient variation to
calibrate the photosynthesis submodel.
10
Traits fANNiiiTraitTraitTraitsAN1....)2,1(maxargmingmaxglSthlSrSshootKrootKThe CO2 response data is shown in Fig. 4(a-b) and the PAR-light-response data in Fig. 4(c-d).
The entire dataset was used to calibrate the C4 photosynthetic parameters using an optimization
algorithm.
Location/weather/management conditions: We considered a drought-prone environment
condition, i.e. Davis, CA in 2010 June-July (see S1) with clay soil (see Table 1) and irrigation
frequency of seven days.
Method to implement crop design framework
There are several approaches to solve this optimization problem. Here, we utilize a gradient free,
evolutionary optimization strategy. This strategy is selected because, as Figure 5 shows, the cost
functional (Equation 26, when varying only two traits) is non-convex and corrugated. This
highly-corrugated surface has many local maximum. This precludes the utilization of gradient-
based methods, and instead suggests the applicability of stochastic, multistart methods that can
explore the phase space efficiently. We specifically use a genetic algorithm (GA) (a gradient-free
meta-heuristic evolutionary search algorithm) to identify the optimal traits. GA is well suited to
multi-modal, highly corrugated solution spaces, especially when the cost function is not easily
adapted to gradient-based methods.
Because GA deploys a population of potential solutions distributed over the design space, they
are less prone to getting stuck in shallow local minima. GA is an inherently stochastic method, so
we repeat each optimization multiple times (10 times) to consider statistical significance of
results and attempt to reliably explore the phase space. The implementation framework can be
found in Figure 6.
Results and discussion
Plant Physiology Model Validation
The physiological model is implemented in MATLAB with inputs of soil and hourly weather
data over a 60-day period. Each model evaluation for a given trait configuration – producing
hourly outputs – took about 40 seconds on a standard laptop.
The plant physiology model builds on the water transport model, and the temperature model
depends on the conservation principles, which are inherently satisfied in our method. Therefore,
we perform a validation exercise on the photosynthetic submodel. The excerise was performed
where marked leaves on 14 maize plants, growing adjacent to the calibration plants, were
monitored repeatedly using the LI-COR 6400 gas exchange system for a day. During that period
the plants were subjected to a diurnal gradient of low to high ambient temperature, and a range of
light. A subset of the plants also had water withheld to evaluate the photosynthetic submodel's
performance under water stress.
11
The photosynthesis submodel, trained on the light, temperature and CO2 response curves,
successfully predicted the photosynthesis of the 14 validation plants during the day of drought
and varying temperature (Fig. 7). Current models of photosynthesis do not account for major
damage to the photosynthetic apparatus in a mechanistic manner. Thus, the model is unable to
predict photosynthetic rates of a couple of points that represented very severely stressed plants.
Design of ideotypes
We deployed the crop design framework on the computing clusters available at Iowa State
(CyEnce cluster) and via NSF XSEDE resources at TACC (Stampede). The simulations usually
took about 4 hours to run for each optimization run on a server with 16 core 2.0 GHz Processer
with 128 GB RAM. Optimizations were initialized with different random seeds and rerun 10
times. In this process, over one million distinct trait combinations were evaluated, and 10
ideotypes were designed.
Comparison between Designed Crop and a Typical Crop
Figure 8 (A) shows that the ideotype produces 10% higher net photosynthesis (yield) than that of
the typical maize. To investigate that, we compare and explore the performances of those two
crops on the hottest day of the season, June 27. The weather on that day is shown in Figure 8,
with an average daytime temperature of 39.810C. Starting from the morning, the hourly value of
solar radiation increases till midday and then decreases till sunset. Relative humidity is high at
night and it decreases during the daytime. Hourly precipitations on that day are zero. It is noted
that the atmospheric temperature increased as the day progresses and went as high as 430C. This
temperature is higher than the optimum functional temperature of maize plants. Thus, plants that
can cool their leaves are desirable.
Figure 8 (D-E) reveals that for the above weather inputs, at the early and later part of the day,
there is no significant variation of hourly net Photosynthesis (An) between the two plants.
However, significant variation is observed at the midday. Photosynthesis depends on CO2, PAR,
enzyme and temperature.
Midday generally has enough PAR, so An depends on CO2 supply, enzyme performance and
temperature. In our study, the enzyme performance profile is the same for both of the plants.
Therefore, midday variation of An depends on the supply of CO2 and temperature.
Figure 8 shows that during the midday period the average CO2 supply is 180 and 200 ppm for
the typical maize and ideotype, respectively. The leaf temperature of the ideotype is lower than
that of the typical maize. These two conditions enable a higher An for the ideotype (Fig 8). The
lower temperature facilitates a shorter duration in which the leaf temperature is above the
optimum temperature for photosynthesis.
Reduced CO2 and lower temperature are related to high stomatal conductance, as shown in Fig.
8. Therefore, Figure 8 (F &G) shows that the typical maize plant has lower stomatal conductance
than that of an ideotype, but should create higher concentration gradient by lower CO2
12
concentration than that of an ideotype in the early part of the day. Figure 8 (E & H) shows that,
due to low stomatal conductance of the typical plant, the cooling of the leaf due to transpiration
of water is lower, and as a result, leaf temperature is higher than that of an ideotype. In short, the
main driver behind the increase of photosynthesis for the ideotype is the positive shift of
stomatal opening operating range.
Values and Significance of the Traits of the optimized ideotype
The positive shift of stomatal conductance depends on the convoluted effect of the seven traits.
Here, we present and discuss on the values and significance of the traits.
Designed Values of the Traits
Our framework provides ten different combinations of traits for the designed crop (Fig 9), all of
them has the same yield (photosynthesis). Based on the variability, we came up with following
two hypotheses:
1) any values of traits within the upper and lower values of suggested traits can be a design crop;
2) some traits might be insensitive for our condition and some traits have threshold values after
which they are insensitive.
Hypothesis 1 does not hold when we use arbitrary combinations of traits within the ranges (see
Fig S2). To test the second hypothesis, we perform sensitivity analysis for all ten combinations
by varying one parameter within the allowable range while keeping the others fixed. The
sensitivity plots indicate that there are threshold values for all of the parameters. There are higher
limits for Sl, ψth, Sr and lower limits for gmin, gmax, kshoot and kroot. Maintaining those threshold
values and subtly changing the other threshold values in the opposite direction shows the
decrease in photosynthesis.
Significance of the Traits
In short, photosynthesis mostly depends on the available sunlight, CO2 and temperature. Among
these parameters, CO2 and temperature can be optimized by adjusting hydraulic parameters.
Among the hydraulic parameters: gmin and gmax are the most directly constraining. gst (a direct
function of gmin and gmax) is the dominating parameter in the transport pathways (causes the
highest path resistance for both CO2 and H2O transports). Rate of water transpiration controls
the cooling effect on the leaves. Cooling (not freezing) is always beneficial during the night (it
reduces the cost of respiration), however, during day it may have positive or negative effect on
the photosynthesis based on the optimum temperature for the enzyme activity (Fig 8). Other
variables including kshoot, Sl, Sr, ψth, kroot primarily ensure that the plant does not reach the
permanent wilting potential. These values will affect the gst if the plant senses water scarcity
when leaf water potential reaches the 'red alert' point, indicated by ψth value.
More specifically, for our designed ideotype
13
• gmax should be higher than a specific value, so that the plant is able to use its available
photosynthetic capacity. The required value of gmax mostly depends on the highest solar
radiation (PPF: photosensitive photon flux density, mostly 400 and 700 nm).
• gmin should be as high as possible to reduce night-time respiration cost, although it may
increase the irrigation cost for the season. gmin sets the minimum transpiration the plant
can do under high VPD or extreme soil water deficit, and thus affects the rate of water
depletion under the most extreme of circumstances.
• Our designed ideotypes do not feel water stress (i.e. red alert) in well-watered conditions,
and are never forced to adjust gst due to water-related issues. However, an ideotype's
kplant and red alert value, ψth, must be selected appropriately. A lower kplant than the
specified value may cause a leaf water potential that is lower than the 'red alert' value,
i.e. the plant will register water stress. A similar effect will happen if the plant increases
(less negative) the ψth value. Required values for kplant and ψth have an inverse
relationship. Therefore, to reduce the cost of root generation, the plant should operate at
the lowest (more negative) possible ψth, i.e. a value close to the permanent wilting
potential, thus lowering the kplant requirement, i.e. it refers to less root production.
Tradeoff between below ground mass and photosynthesis/above ground mass
Next, we extrapolate our results from our design space (reproductive stage) to vegetative stage.
During the vegetative stage, to increase the photosynthetic capability of the plant (LAI), a bigger
shoots could be better, whereas smaller roots would help the plant to invest more resources into
growing the shoot. Therefore, we explore the effect of a smaller root, i.e. smaller kroot, on the net
photosynthesis of the ideotypes.
Figure 10 shows that a 50% reduction of kroot from 40 to 20 reduces the net photosynthesis by
only 0.06%. For the reduced kroot plant, the leaf water potential sometimes reaches lower than the
threshold leaf water potential (ψth) and closes (Figure 11). The modified plant has the potential to
increase LAI during the vegetative stage, leading to improvement in net photosynthesis (yield).
Conclusions
An integrated framework of optimization, thermodynamics and plant physiology was developed
to design a crop ideotype. The backbone of the framework is a 1-D plant physiology model and
the coupling of transport and energy conversion models based on laws of thermodynamics. The
models were augmented with a nature-inspired meta-heuristic optimization method, the genetic
algorithm (GA), and was implemented via the MATLAB® software. The framework was used to
design maize crop for a drought-prone weather condition in Davis, CA. Seven physiological
traits which are primarily related to plant hydraulics and ultimately affect the photosynthesis and
water usage were considered in the study. The traits are minimum stomatal opening (
);
maximum stomatal opening (
); sensitivity of stomatal opening with leaf water potential (
); threshold bulk leaf water potential at stomatal closure (
); and sensitivity of
with root
14
mingmaxglSthlSwater potential (
); shoot hydraulic conductance (
); and root hydraulic conductance (
). With enough irrigation, the designed crop showed 10% improvement in yield, and
,
and
are found to be the vital traits. Currently, the model is using hourly data; however,
it could be easily modified for more frequent data. The framework is modular and can be easily
augmented with other existing mechanistic models to capture more physics. The developed tool
can help plant breeders and scientists to determine the optimal crop ideotypes for various
climates (climate-smart crops) and locations. Integration of the developed framework with
breeding programs can speed the crop development process, wherein the framework can be used
to propose ideotypes for target environments and the breeder can breed plants like based on the
ideotypes.
Potentially, ideotypes designed using a different crop model might look different from those
presented, as shown for wheat in [40]. These models can simulate observed yields under a range
of environments for the current conditions. However, simulated climate change impacts could
vary across models due to differences in model structures and parameter values [41]. Further
improvements of crop models and a more rigorous framework will be required for robust crop
ideotype design.
ACKNOWLEDGEMENTS
T.Z.J, B.G, D.A gratefully acknowledge financial support from the Presidential initiative for
interdisciplinary research of Iowa State University. B.G and T.Z.J gratefully acknowledge the
Plant Science Institute at Iowa State University, and computing support via NSF XSEDE
CTS110007. M.E.G gratefully acknowledges support via USDA National Institute of Food and
Agriculture, Hatch project number #1001480.
AUTHOR CONTRIBUTIONS:
D.A formed the interdisciplinary team and proposed to couple numerical optimization and plant
physiology, D.A, M.E.G, B.G designed research plan, M.E.G. designed the crop hydraulic
model, B.G. implemented model, all authors improved the model, B.G and T.Z.J designed the
optimization framework, T.Z.J ran the simulations and processed the data, all authors analyzed
data and wrote the paper.
COMPETING FINANCIAL INTERESTS
The author(s) declare no competing financial interests.
15
rSshootKrootKmaxgmingthReferences
[1] H. C. J. Godfray et al., "Food Security: The Challenge of Feeding 9 Billion People,"
Science, vol. 327, no. 5967, pp. 812–818, Feb. 2010.
[2] A. J. Challinor, F. Ewert, S. Arnold, E. Simelton, and E. Fraser, "Crops and climate change:
progress, trends, and challenges in simulating impacts and informing adaptation," J. Exp.
Bot., vol. 60, no. 10, pp. 2775–2789, Jul. 2009.
[3] K. J. Boote, M. J. Kropff, and P. S. Bindraban, "Physiology and modelling of traits in crop
plants: implications for genetic improvement," Agric. Syst., vol. 70, no. 2–3, pp. 395–420,
Nov. 2001.
[4] P. Martre, B. Quilot-Turion, D. Luquet, M.-M. O.-S. Memmah, K. Chenu, and P. Debaeke,
"Model-assisted phenotyping and ideotype design," in Crop Physiology, Elsevier, 2015, pp.
349–373.
[5] K. H. Jensen, J. Lee, T. Bohr, H. Bruus, N. M. Holbrook, and M. A. Zwieniecki,
"Optimality of the Münch mechanism for translocation of sugars in plants," J. R. Soc.
Interface, vol. 8, no. 61, pp. 1155–1165, Aug. 2011.
[6] C. Doussan, L. Pagès, and G. Vercambre, "Modelling of the Hydraulic Architecture of Root
Systems: An Integrated Approach to Water Absorption-Model Description," Ann. Bot.,
vol. 81, no. 2, pp. 213–223, Feb. 1998.
[7] C.-T. Lai and G. Katul, "The dynamic role of root-water uptake in coupling potential to
actual transpiration," Adv. Water Resour., vol. 23, no. 4, pp. 427–439, Jan. 2000.
[8] P. J. Sellers, Y. Mintz, Y. C. Sud, and A. Dalcher, "A Simple Biosphere Model (SIB) for
Use within General Circulation Models," J. Atmospheric Sci., vol. 43, no. 6, pp. 505–531,
Mar. 1986.
[9] T. Vogel, M. Dohnal, J. Dusek, J. Votrubova, and M. Tesar, "Macroscopic Modeling of
Plant Water Uptake in a Forest Stand Involving Root-Mediated Soil Water Redistribution,"
Vadose Zone J., vol. 12, no. 1, Feb. 2013.
[10] J. Rings et al., "Bayesian Inference of Tree Water Relations Using a Soil-Tree-Atmosphere
Continuum Model," Procedia Environ. Sci., vol. 19, pp. 26–36, Jan. 2013.
[11] V. Couvreur, J. Vanderborght, and M. Javaux, "A simple three-dimensional macroscopic
root water uptake model based on the hydraulic architecture approach," Hydrol Earth Syst
Sci, vol. 16, no. 8, pp. 2957–2971, Aug. 2012.
[12] J. Wang, Q. Yu, and X. Lee, "Simulation of crop growth and energy and carbon dioxide
fluxes at different time steps from hourly to daily," Hydrol. Process., vol. 21, no. 18, pp.
2474–2492, Aug. 2007.
[13] K. Steppe, D. J. W. De Pauw, R. Lemeur, and P. A. Vanrolleghem, "A mathematical model
linking tree sap flow dynamics to daily stem diameter fluctuations and radial stem growth,"
Tree Physiol., vol. 26, no. 3, pp. 257–273, Mar. 2006.
[14] F. Somma, J. W. Hopmans, and V. Clausnitzer, "Transient three-dimensional modeling of
soil water and solute transport with simultaneous root growth, root water and nutrient
uptake," Plant Soil, vol. 202, no. 2, pp. 281–293, May 1998.
[15] R. A. Duursma and B. E. Medlyn, "MAESPA: a model to study interactions between water
limitation, environmental drivers and vegetation function at tree and stand levels, with an
example application to [CO2] ? drought interactions," Geosci. Model Dev. Katlenburg-
Lindau, vol. 5, no. 4, p. 919, 2012.
[16] T. H. van den Honert, "Water transport in plants as a catenary process," Discuss. Faraday
Soc., vol. 3, no. 0, pp. 146–153, 1948.
16
[17] A. Askarzadeh, "Bird mating optimizer: An optimization algorithm inspired by bird mating
strategies," Commun. Nonlinear Sci. Numer. Simul., vol. 19, no. 4, pp. 1213–1228, Apr.
2014.
[18] I. F. Jr, X.-S. Yang, D. Fister, and I. Fister, "Cuckoo Search: A Brief Literature Review," in
Cuckoo Search and Firefly Algorithm, X.-S. Yang, Ed. Springer International Publishing,
2014, pp. 49–62.
[19] K. Elofsson, G. Bengtsson, and I.-M. Gren, "Optimal Management of Invasive Species with
Different Reproduction and Survival Strategies," Nat. Resour. Model., vol. 25, no. 4, pp.
599–628, Nov. 2012.
[20] C. Riedelsheimer and A. E. Melchinger, "Optimizing the allocation of resources for
genomic selection in one breeding cycle," TAG Theor. Appl. Genet. Theor. Angew. Genet.,
vol. 126, no. 11, pp. 2835–2848, Nov. 2013.
[21] C. XiangWei, Y. TianLi, and Y. GongMing, "Research advances in non-destructive
prediction technologies using VIS/NIR spectroscopy for kiwifruit property," Transactions
of the Chinese Society of Agricultural Engineering, 20-Jul-2006. [Online]. Available:
https://eurekamag.com/research/012/927/012927668.php. [Accessed: 21-Mar-2017].
[22] G. Decoux and F. Hospital, "Popmin: A Program for the Numerical Optimization of
Population Sizes in Marker-Assisted Backcross Programs," J. Hered., vol. 93, no. 5, pp.
383–384, Sep. 2002.
[23] S. Manzoni, G. Vico, S. Palmroth, A. Porporato, and G. Katul, "Optimization of stomatal
conductance for maximum carbon gain under dynamic soil moisture," Adv. Water Resour.,
vol. 62, pp. 90–105, Dec. 2013.
[24] M. A. Semenov and P. Stratonovitch, "Designing high-yielding wheat ideotypes for a
changing climate," Food Energy Secur., vol. 2, no. 3, pp. 185–196, Dec. 2013.
[25] V. Picheny et al., "Finding realistic and efficient plant phenotypes using numerical
models," ArXiv Prepr. ArXiv160303238, 2016.
[26] N. Park, Physicochemical & environmental plant physiology, Park vols. Academic Press,
2009.
[27] I. R. Cowan, "Transport of Water in the Soil-Plant-Atmosphere System," J. Appl. Ecol.,
vol. 2, no. 1, pp. 221–239, 1965.
[28] G. S. Campbell and J. M. Norman, An introduction to environmental biophysics. New
York, NY, USA: Springer, 1998.
[29] A. L. Buck, "New Equations for Computing Vapor Pressure and Enhancement Factor," J.
Appl. Meteorol., vol. 20, no. 12, pp. 1527–1532, Dec. 1981.
[30] S. V. Caemmerer, Biochemical models of leaf photosynthesis. 2000.
[31] G. J. Collatz, M. Ribas-Carbo, and J. A. Berry, "Coupled Photosynthesis-Stomatal
Conductance Model for Leaves of C4 Plants," Funct. Plant Biol., vol. 19, no. 5, pp. 519–
538, Oct. 1992.
[32] R. G. Allen, L. S. Pereira, D. Raes, M. Smith, and others, "Crop evapotranspiration-
Guidelines for computing crop water requirements-FAO Irrigation and drainage paper 56,"
FAO Rome, vol. 300, no. 9, p. D05109, 1998.
[33] Z. Dong, O. Danilevskaya, T. Abadie, C. Messina, N. Coles, and M. Cooper, "A Gene
Regulatory Network Model for Floral Transition of the Shoot Apex in Maize and Its
Dynamic Modeling," PLOS ONE, vol. 7, no. 8, p. e43450, Aug. 2012.
17
[34] M. A. Zwieniecki, H. A. Stone, A. Leigh, C. K. Boyce, and N. M. Holbrook, "Hydraulic
design of pine needles: one-dimensional optimization for single-vein leaves," Plant Cell
Environ., vol. 29, no. 5, pp. 803–809, May 2006.
[35] D. N. L. Menge, F. Ballantyne, and J. S. Weitz, "Dynamics of nutrient uptake strategies:
lessons from the tortoise and the hare," Theor. Ecol., vol. 4, no. 2, pp. 163–177, May 2011.
[36] V. M. Dunbabin, S. McDermott, and A. G. Bengough, "Upscaling from Rhizosphere to
Whole Root System: Modelling the Effects of Phospholipid Surfactants on Water and
Nutrient Uptake," Plant Soil, vol. 283, no. 1–2, pp. 57–72, May 2006.
[37] A. J. Guswa, "Soil-moisture limits on plant uptake: An upscaled relationship for water-
limited ecosystems," Adv. Water Resour., vol. 28, no. 6, pp. 543–552, Jun. 2005.
[38] G. Damour, T. Simonneau, H. Cochard, and L. Urban, "An overview of models of stomatal
conductance at the leaf level," Plant Cell Environ., vol. 33, no. 9, pp. 1419–1438, Sep.
2010.
[39] T. J. Brodribb and N. M. Holbrook, "Declining hydraulic efficiency as transpiring leaves
desiccate: two types of response*," Plant Cell Environ., vol. 29, no. 12, pp. 2205–2215,
Dec. 2006.
[40] S. Asseng et al., "Uncertainty in simulating wheat yields under climate change," Nat. Clim.
Change, vol. 3, no. 9, pp. 827–832, Sep. 2013.
[41] R. P. Rötter, T. R. Carter, J. E. Olesen, and J. R. Porter, "Crop-climate models need an
overhaul," Nat. Clim. Change Lond., vol. 1, no. 4, pp. 175–177, Jul. 2011.
18
Figures and Legends
Figure 1: A conventional resistance or conductance (resistance=1/conductance) model of maize
hydraulics (panel a), and the model used for simulating maize hydraulics including feedbacks
(panel b). The conductance of the stomata to water vapor (gst) and CO2 is modulated by the
water potential of the leaf (ψleaf) if below a threshold (ψth). The stomatal opening is scaled to the
maximum stomatal conductance (gmax; a proxy of how many stomata and how wide they open)
which sets the maximum water loss rate and the maximum CO2 uptake rate for sunlit leaves.
How effective stomatal closure is (minimum stomatal conductance; gmin) is determined by
cuticle waxes which stop water loss from the leaf surface, affecting the rate of desiccation under
drought, but this state also prevents CO2 uptake. The slope of the response is tuned by an
inherent sensitivity (Sl) or a contribution of the root, based upon sensing of soil drying (Sr). The
supply of water is proportional to the difference of water potential between soil and air, and
inversely proportional to three conductances in series: Ksoil, Kroot, and Kstem+leaf. The
demand for water is driven by environmental variables: the boundary layer conductance and the
temperature of the leaf. The leaf temperature is determined through an energy balance and
influences both transpiration and a coupled model of photosynthesis.
19
ψsoilRsoilRrootRstemRleafTrrstrblcψleafψair1/KsoilYsoilYleaf1/Kleaf1/Kstem1/Kroot1/gblcYairTr1/gsta1/Kroot1/gblcψleaf ψth ψsoilTr1/gmaxbreaks at ψleaf <ψth 1/gmin-1/gmaxResistanceadjustment1/gst1/gstΨair ψrootf(Sl ,Sr)1/Kshoot1/gst-1/gmaxbFigure 1Figure 2: Model input, output and connectivity among the submodels.
20
Figure 3: Flowchart for the implementation of plant physiology model
21
Evaluate ψleaf by satisfying supply and demand of water via minimizing (Jw,d-Jw,s)2. Supply of water from soil to the leaf, Jw,d, and evaporation of water vapor from leaf to the atmosphere Jw,s, are calculated from equations, (1) and (6), respectively. Evaluate leaf temperature (Tleaf) by considering energy equilibrium on the leaf via satisfying equation (23)Evaluate CO2concentration inside the intercellular space of the leaf (Cc,i) by satisfying transport balance via minimizing (JC,s-JC,d)2. supply of CO2 from the environment, JC,sand the damend of CO2by the plant, JC,d, are calculated from equations (10) and (22), respectively.At a particular time step, ticalculate soil water potential using Eq. 4Guess stomatal conductance, gst(gst,new–gst)2<tolCalculate corrected gst,newusing Eq25Two factors must be favorable for a leaf to remain alive: leaf water potential (ψwilt) and temperature (Tdamage) must remain within nonlethal bounds. Check for survival of plant: if ψleaf ≤ψwilt or Tleaf ≤TdamageUsing corrected gst,new, evaluate ψleaf, Tleaf, and Cc,iusing steps 3,4 and 5, respectivley Store water loss/usage, Tr = Jw,d, and net CO2consumption, An = Jc,das photoCumulative photosynthesis, Photo = sum(photoiΔt), and cumulative water usage =sum(TriΔt) gst = gst,newti+1= ti+Δt, Δt is the period of the inputted weather dataAverage net photosynthesis must be positive for a leaf to remain alive. Output: Photo and Water UsageOutput:Photo = 0yesnoyesnoyesnonoyesti+1<periodStartFigure S1Figure 4: (a-b) CO2 responses of maize photosynthesis at varying leaf temperatures for two
plants used in the calibration of the photosynthesis submodel. Lines connect points measured at
the same leaf temperature. (c-d) Response of maize photosynthesis to photosynthetically active
radiation (PAR) measured at varying leaf temperatures for two plants used in the calibration.
Lines connect points measured at the same leaf temperature.
22
-1001020304050-1001020304050Predicted An ( mol/m 2/s)Observed An ( mol/m 2/s)abcdeFigure S4
Figure 5: Distribution of photosynthesis, in terms of CO2 assimilation, within the selected
ranges of gmin and ψth for typical maize as in Table 1.
23
Figure 6: Flowchart of crop design framework implementation.
24
Figure 7: Validation of the C4 photosynthesis submodel. The circles represent the predicted and
observed calibration data for light, CO2 and temperature responses for plant 1 and triangles for
plant 2. The diamonds represent the validation data: the observed and predicted photosynthesis
of 14 plants measured over a day varying in drought treatments, temperature and light. The
points that deviate from the 1:1 relationship were plants that underwent the greatest drought
stress during the hottest time of the day, and represent photosynthetic inhibition or damage.
25
-1001020304050-1001020304050Predicted An ( mol/m 2/s)Observed An ( mol/m 2/s)abcdeFigure S4
Figure 8: (A) Hourly cumulative photosynthesis for the typical plant (red line) and the sets of
simulated ideotypes at well-irrigated condition. For simulated plants, the profiles are overlapped.
Here, instead of days of the months, time has been presented as hours. (B-C) Hourly variation of
solar radiation and relative humidity on June 27. (D-E) Hourly variation of photosynthesis (An),
concentration inside the leaf. (F) Green for PEP limited case, Orange for Rubisco
and CO2
limited case, Red symbol (for ideotypes) is the net effect with CO2 concentration 180 ppm, Gray
symbol (for typical) is the net effect with CO2 concentration 120 ppm. Blue lines are light
(PAR) limited case. (G-H) Hourly variation of variation stomatal conductance with respect to
minimum conductance, and leaf temperature.
26
4am8am124pm8pmAn(mol/m2)0102030404am8am124pm8pmgst-gmin(mol/mss)00.10.20.34am8am124pm8pmCi(ppm)02004006004am8am124pm8pmPAR(mmol/m2s)Ta(o)C00.511.5225303540454am8am124pm8pmRH(%)Uc(m/s)406080123June1June15July31Photo(mol/m2)20406080Photo4am8am124pm8pmTleaf(oC)2025303540An(mol/m2s)Tleaf(0C)0204060204011am1pm9am6am7am6pm4pm2pmABCDEFGHFigure 9: Variation of the traits among the outcomes obtained via optimizations. The big red
symbol corresponds to the typical value and the numbers indicate the upper and lower limits of
the traits (Table 1).
27
Figure 10: (top) Effect of kroot on the photosynthesis (mid-bottom) hourly variation of stomatal
conductance and leaf water potential for the designed ideotype (red) and modified ideotype
(blue).
28
Table 1 Management/Agronomic parameters, plant physiological parameters of typical maize,
and bio-chemistry parameters for gas exchange calculations.
Management/Agronomic parameters
, Soil-texture-dependent parameter (unit less)
, Saturated soil hydraulic conductivity [mol m-1 s-1 MPa-1]
[MPa]
( Saturated water content in the soil) [m3 m-3]
(depth of soil) [m]
Rs, Radius of soil occupied/supplied for one plant [m]
Irrigation frequency (day)
Plant physiological parameters
L, Root length density (m m-3)
rroot, Root radius at the end of rhizosphere (m)
d, leaf width (m)
Hc, Height of the plant (m)
Permanent wilting soil water potential for leaf wilting (MPa)
Temperature for permanent leaf damage (0C)
LAI, Leaf area index, [m2m-2]
Plant Hydraulic Parameters
, minimum stomatal conductance mol m-2s-1
, maximum stomatal conductance mol m
-2
-1
s
, response of stomatal conductance with leaf water potential mol m
-2
s
-1
MPa
-1
, response of stomatal conductance upon sensing drying soil mol m
-2
-1
s
MPa
-2
, threshold of leaf water potential for stomatal closure, MPa
, hydraulic conductance of leaf mmol m
-2
s
-1
MPa
-1
Kroot, hydraulic conductance of root mmol m-2s-1MPa-1
Kroot, hydraulic conductance of stem mmol m-2s-1MPa-1
Bio-chemistry parameters from gas exchange data
14.95
1.69
-0.00598
0.39
1
0.1128
7
15200
0.0005
0.1
1
-1.33
60
1
0.02
0.25
15
-1.25
200
45
45
45
(unitless)
(unitless)
0.0036
(unitless)
1.1693 x (unitless)
4.93
0.0844
(unitless)
(unit less)
(unit less) 13.1
(unitless)
0.00445
(unit less)
(unit less)
A (unit less)
0.945
0.094
(unitless)
0.0715
[mol m-2 s-1] 322.3
B (unit less)
31.6
[mol m-2 s-1]
51
[mol m-2 s-1]
[mol m-2 s-1]
1.39
1.55
5.57
2.28
96.1
29
bsatksatsatsHmingmaxglSrSthleafK1gbJmax,10PEPVQ2g10,QdRmax,,10RVPEPQaJmax,10ROVQ25,tR25max,J25,,RPEPV25max,PEPVC (unit less)
42.1
[mol m-2 s-1]
126.7
30
25max,ROVTable 2 Bounds to hydraulic parameters varied in the genetic algorithm.
Lower
bound
Upper
bound
mol m-2s-1
mol m-2s-1MPa-1 mol m-2s-1MPa-2 MPa
mmol m-2s-1 MPa-1
1e-6
0.07
=0
=0
-1.33
1e-6
1e-6
3
3.0
=89.99
=89.99
0
30
60
31
mingmaxgllStanrrStanthstemleafKrootKlrlrNomenclature
-
-
-
Absorptance of leaf
Absorptance of leaf for thermal infrared radiation
Constant related with carboxylation rate
[mol m-2 s-1]
Hourly net assimilation or net photosynthesis
-
-
-
-
-
kPa
-
° C
-
Soil texture dependent parameter
Parameter related with probabilistic GA formulation
Parameter related with probabilistic GA formulation
Parameter related with probabilistic GA formulation
Constant related with carboxylation rate
Constant related with Tetens formula
Constant related with Tetens formula
Constant related with Tetens formula
Constant related with carboxylation rate
[Pa Pa-1] or ppm
Atmosphere CO2 partial pressure, or concentration
[Pa Pa-1] or ppm
Intercellular airspace CO2 partial pressure, or concentration
[Pa Pa-1]
Mesophyll CO2 partial pressure, or concentration.
[J mol-1K-1]
Specific heat of air.
[m]
[m]
[m2s-1]
[m2s-1]
Leaf average width
Parameter related with leaf width
Thermal diffusivity in air
Water vapor diffusivity in air
-
Fraction of PAR contributes to the Photosystem II
32
aIRaAAnb0b1b2bB0c1c2cCacC,icC,mcC,PCdedHDWVDPSIIPARf_
-
-
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
-
[m]
[m]
[m]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
lant physiological parameter related with photosynthesis or
CO2 assimilation
lant physiological parameter related with photosynthesis or
CO2 assimilation
Boundary conductance to water transport
Boundary conductance to heat transfer on leaf surface.
Minimum Stomatal conductance, or stomatal conductance at
light compensation point, minimum stomatal conductance to
water vapor including epidermal conductance
Maximum Stomata conductance
Bundle-sheath conductance to CO2
Effective stomata conductance to water
Grashof number
Depth of soil
Height of the plant
Height at which wind speed obtained from weather data
A parameter related with electron transport rate
Physiological parameter related with electron transport rate
Physiological parameter related with electron transport rate
CO2 demand by photosynthetic activity based enzyme limited
condition
CO2 demand by photosynthetic activity based sunlight limited
condition
total electron transport rate is at leaf temperature
total electron transport rate is at 25 ° C
[mol m-2 s-1]
Maximum total electron transport rate is at 25 ° C
33
1g2gblcghbcgmingmaxgbsCg,stgGrsHcHmHIaJbJ(enzyme),dCJ)(,lightJdCteJ,25,eJ25max,eJ
[mol m-2 s-1]
[mol m-2 s-1]
Rate of water vapor demanded by atmosphere from leaf
Rate of water supplied from soil to leaf
[mol m-2 s-1 Pa-1]
hydraulic conductivity of saturated soil
[mol m-2 s-1 Pa-1]
leaf hydraulic conductance to water
[mol m-2 s-1 Pa-1]
the Michaelis-Menten constant
[mol m-2 s-1 Pa-1]
plant hydraulic conductance to water
[mol m-2 s-1 Pa-1]
root hydraulic conductance to water
[mol m-2 s-1 Pa-1]
soil hydraulic conductance to water
[mol m-2 s-1 Pa-1]
stem hydraulic conductance to water
[m m-3]
[J mol-1]
[m2m-2]
-
-
[mol m-2 s-1]
[mol m-2 ]
[mol m-2 s-1]
-
[Pa]
[Pa]
[Pa]
Root length density , root length per unit volume of soil
Latent heat of vaporization of water.
Leaf area index
a factor related with zero plane displacement for wind speed
a factor related to the momentum roughness parameter for
wind speed
Hourly net CO2 assimilation
Total photo
Photo active radiation
Initial population in genetic algorithm
Atmospheric pressure
Prandtl number
Vapor pressure of air
Vapor pressure of leaf surface
34
dWJ,sWJ,satkleafKpKplantKrootKsoilKstemKLvapLLAImnphotoPhotoPARoPaPPrvaPvlP
-
-
-
-
-
Q10 coefficient conversion factor related to mitochondrial
respiration calculation
Q10 coefficient conversion factor related to maximum PEP
carboxylation
Q10 coefficient conversion factor related to maximum PEP
regeneration
Q10 coefficient conversion factor related to maximum rubisco
carboxylation
Reflectance, i.e. amount of sunlight reflected from
surroundings,
the
[m]
Root radius including rhizosphere
[mol m-2 s-1]
[mol m-2 s-1]
-
-
[W m -2]
-
Total mitochondrial respiration in the mesophyll and bundle
sheath at leaf temperature
Total mitochondrial respiration in the mesophyll and bundle
sheath at 25° C temperature
Reynolds number
Relative humidity of surrounding air.
Solar radiation,
Schmidt number
[mol m-2 s-1MPa-1]
Slope of stomatal conductance with leaf water potential
[mol m-2 s-1MPa-2]
Slope of
with root water potential
[° C]
[° C]
[° C]
[mol m-2 s-1]
[m s-1]
[m s-1]
Temperature of air
Temperature of leaf
leaf temperature at permanent leaf damage
Rate of water transpires from soil to environment through the
plant
Wind speed on the canopy
Wind speed from weather data
35
RtQ,10max,10PEPVQmax,,10RVPEPQmax,10ROVQrrootrtR25,tRReRHScSlSrSlSaTleafTwiltTTrcUmU
[m s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[mol m-2 s-1]
[Pa Pa-1]
Shearing velocity
Effective Rate of PEP carboxylation at leaf temperature given
by Michaelis-Menten Equation
Maximum PEP carboxylation rate at leaf temperature
Maximum PEP carboxylation rate at 25 C temperature
PEP regeneration rate at leaf temperature
PEP regeneration rate at 25 C temperature
Maximum rubisco carboxylation rate at leaf temperature
Maximum rubisco carboxylation rate at 25 C temperature
Vapor pressure deficit between
atmosphere
intercellular space and
[mol m-2 s-1]
Transpiration of water through the plant
Greek
-
-
-
-
-
-
is a fitting parameter related to photosynthesis rate
parameter to make the exponent dimensionless in stomatal
conductance model
empirical parameter related with boundary layer conductance
ratio of CO2 conductance and water vapor conductance through
stomata
Empirical curvature factor related with electron transport rate
ratio of CO2 conductance and water vapor conductance through
air boundary layer
[m2s-1]
[m3m-3]
[m3m-3]
kinetic viscosity of air
Volumetric water content in the soil
Soil volumetric saturation
36
*UPEPVmaxPEPV25max,PEPVRPEPV,25,,RPEPVmaxROV25max,ROVVPDLossWater xZasat
[Pa]
[Pa]
[Pa]
[Pa]
[Pa]
[Pa]
Leaf water potential
Root water potential
Saturated soil water potential
Soil water potential
Soil water potential at root
Threshold leaf water potential at stomatal closure
[W m−2 K−4]
Stefan Boltzmann constant
37
leafrootsatsoilrootsoil,thSupplementary Materials
Integrating optimization with thermodynamics and plant physiology for crop
ideotype design
Talukder Z. Jubery1, Baskar Ganapathysubramanian1, Matthew E. Gilbert2, and Daniel
Attinger1
1Department of Mechanical Engineering and Department of Electrical and Computer
Engineering, Iowa State University, Ames, IA, 50011;
2Department of Plant Sciences, University of California, Davis, CA 95616, USA.
*Correspondence to: [email protected], [email protected], [email protected]
Justification for selecting June-July weather
Maximum solar radiance is observed in this period of year in the northern hemisphere. Our
hypothesis was that a crop should be the most productive in this period provided that the plant
has access to adequate water and nutrients. Due to high solar radiance that results high
temperature, and generally low relative humidity in drought-prone areas, this period should also
mimic the highest water demand by the environment from the plant.
We considered that the typical maize plant was fully grown, i.e. at the beginning of full canopy
closure, and total yielding period was two months. During this two-month period average solar
radiation was 489.17 W/m2, relative humidity 45.63%, air temperature was 30.6 0C, and wind
speed was 3.048 m/s, precipitation does not occur in drought conditions.. The hourly variations
of weather parameters are in Figure S1.
38
Fig. S1 Hourly variation of solar radiation (S), relative humidity (RH), atmospheric temperature
(Ta), wind speed (Um), precipitation (Precip) and photosynthetically active radiation (PAR) in
June-July 2010, Davis, CA.
39
HoursUm(m/s)050010000246810HoursPrecip(mm)0500100000.511.5HoursPAR(W/m2)050010000500100015002000HoursS(kW/m2)0500100000.20.40.60.81HoursRH(%)0500100020406080HoursTa(oC)0500100010203040ABCDEFFigure S2Figure S2: Sensitivity analysis of the traits obtained from GA optimization.
40
|
1605.08314 | 1 | 1605 | 2016-05-26T14:53:04 | Statistical mechanics of the Huxley-Simmons model | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech"
] | The chemomechanical model of Huxley and Simmons (HS) [A. F. Huxley and R. M. Simmons, Nature 233, 533 (1971)] provides a paradigmatic description of mechanically induced collective conformational changes relevant in a variety of biological contexts, from muscles power-stroke and hair cell gating to integrin binding and hairpin unzipping. We develop a statistical mechanical perspective on the HS model by exploiting a formal analogy with a paramagnetic Ising model. We first study the equilibrium HS model with a finite number of elements and compute explicitly its mechanical and thermal properties. To model kinetics, we derive a master equation and solve it for several loading protocols. The developed formalism is applicable to a broad range of allosteric systems with mean-field interactions. | physics.bio-ph | physics |
Statistical mechanics of the Huxley-Simmons model
M. Caruel1, ∗ and L. Truskinovsky2
1MSME, CNRS-UMR 8208, 61 Avenue du G´en´eral de Gaulle, 94010 Cr´eteil, France
2LMS, CNRS-UMR 7649, Ecole Polytechnique, 91128 Palaiseau Cedex, France
(Dated: May 27, 2016)
The chemomechanical model of Huxley and Simmons (HS) [A. F. Huxley and R. M. Simmons,
Nature 233, 533 (1971)] provides a paradigmatic description of mechanically induced collective
conformational changes relevant in a variety of biological contexts, from muscles power-stroke and
hair cell gating to integrin binding and hairpin unzipping. We develop a statistical mechanical
perspective on the HS model by exploiting a formal analogy with a paramagnetic Ising model. We
first study the equilibrium HS model with a finite number of elements and compute explicitly its
mechanical and thermal properties. To model kinetics, we derive a master equation and solve it
for several loading protocols. The developed formalism is applicable to a broad range of allosteric
systems with mean-field interactions.
I.
INTRODUCTION
Passive, mechanically induced conformational change
in a parallel bundle of bistable elements subjected to fi-
nite temperature was first studied theoretically in the
pioneering paper of A.F. Huxley and R.M. Simmons [1].
They modeled in this way the mechanism of fast force
recovery in skeletal muscles subjected to shortening un-
der a length clamp (isometric) protocol. Such loading
defines a hard device ensemble, which has to be distin-
guished from a soft device (isotonic) ensemble exhibiting
some rather different properties [2].
The HS model interpreted the conformational change,
appearing in the muscle context under the name of a
power stroke, in a highly simplified way, as a "digital"
switch between an extended and a contracted states. HS
assumed that the contracted state is biased by the im-
posed shortening and treated the ensuing collective fold-
ing as a deterministic chemical reaction. The information
about the energetic preference of the contracted state and
about the corresponding energy barriers was encoded into
the reaction rates which became functions of the "me-
chanical configuration" of the system.
In the muscle literature the chemomechanical descrip-
tion of HS was later refined through the inclusion of
numerous additional chemical reactions between various
intermediate configurations and their kinetics was mod-
elled phenomenologically [3–7]. Almost identical descrip-
tions of mechanically driven conformational changes were
proposed independently in the studies of cell adhesion
[8, 9] and in the context of hair cell gating [10, 11].
Other closely related systems include mechanical denat-
uration of RNA (ribonucleic acid) and DNA (deoxyri-
bonucleic acid) hairpins [12–14], unzipping of biological
macromolecules [15–21], collective action of SNARE (sol-
uble N-ethylmaleimide sensitive receptor) proteins dur-
ing opening of synaptic pores [22] and even formation
of ripples in graphene sheets [23]. For all these systems
∗ [email protected]
the HS model can be viewed as a fundamental mean-field
prototype.
The goal of the present paper is to reassess the chem-
ical reaction based approach of HS from the perspective
of statistical mechanics for a system with a finite num-
ber of elements while emphasizing the role of fluctua-
tions. In such reformulation of the HS model we follow
the pioneering work of T.L. Hill [3, 24] and more recent
developments in Refs. [2, 25]. The zero temperature HS
model was studied from this viewpoint in Ref. [26] where
it was presented as a version of a fiber bundle model [27].
Here we extend the analysis of Ref. [26] to finite temper-
atures focusing on thermomechanical coupling that has
not been previously addressed in the chemomechanical
framework.
Viewed from an abstract statistical mechanics perspec-
tive, the HS model is quite similar to a paramagnetic
Ising model whose thermodynamic and kinetic proper-
ties are well known [28]. The equivalence, however, is
not complete due to the presence in the HS model of an
elastic spring, buttressing each spin element. Another
complication is the length clamp control which is uncon-
ventional for magnetic analogs of the HS system. Among
the new effects revealed by the HS model, which would be
unusual for paramagnets in an external field, it is enough
to mention negative susceptibility and pseudo-critical be-
havior without genuine cooperativity.
Given that the explicit formulas for the equilibrium
free energy of a spin system with mean-field interactions
are rather straightforward, we can easily access both me-
chanical and thermal properties of the HS model includ-
ing the heat release associated with mechanical loading.
We can also specify the entropic contributions to mechan-
ical and thermal susceptibilities and distinguish adiabatic
from isothermal responses.
To complement the equilibrium picture, we study in
this paper the stochastic dynamics of a HS system with
a finite number of elements. The starting point here is a
thermally induced random walk in the energy landscape
biased by the mechanical loading [29, 30]. We show that
due to the mean-field nature of the interactions, the ki-
netic properties of the HS system are fully determined
2
the HS model allows us to interpret x as a spin vari-
able. The soft spins (snap-springs) version of the HS
model, corresponding to the case when each of the two
energy wells is represented by a quadratic potential, was
developed in [2, 25], however, the comparison of the two
models shows that the additional effects due to elasticity
of the conformational states are of mostly quantitative
nature.
We choose a, denoted by h in Ref. [1], as the "refer-
ence" size of the conformational change equal to the dis-
tance between two infinitely localized energy wells, and
we denote by v0 the intrinsic energy bias distinguishing
the two states; see Fig. 1. The energy of the spin element
can be now written as
v0
vHS(x) =
if x = 0,
if x = −a,
0
∞ otherwise.
The energy v0 is an implicit representation of the ATP-
fueled activity in this otherwise passive system. The pres-
ence of such bias ensures that in the reference state the se-
ries spring is stretched and generates (active or tetanized)
tension.
It will be convenient to use dimensionless variables and
we choose a as our characteristic distance, assuming that
the non-dimensional spin variable x takes values 0 or −1.
We normalize the total energy of the system by κ0a2 and
obtain
(y − x)2,
1
2
v(x; y) = (1 + x)v0 +
(1)
where x = {0,−1} and y is the length of the combined
element that includes a bistable unit and a linear spring.
If we define y0 = v0 − 1/2 and use the muscle mechanics
jargon, we can say that for y > y0 (respectively y <
y0) the global minimum of the energy (1) corresponds
to the pre-power-stroke state (respectively post-power-
stroke state); in [1], the shifted elongation y − y0 was
denoted by y.
Note that our variable y plays a role of the external
(magnetic) field for the spin variable x and therefore
our model resembles the zero dimensional Ising model
of paramagnetism [28]. However, due to the presence of
a linear spring this Ising model is unusual: The external
field has its own "energy" represented by the quadratic
term in y. In the original HS experiments a muscle was
loaded in a hard device which apparently makes this
"energy" irrelevant. However, as we show below, the
quadratic term in y brings additional stiffness into the
overall mechanical response of the system and is there-
fore responsible for some interesting effects.
a. Thermal equilibrium.
Denoting by T the abso-
lute temperature we can write the equilibrium probabil-
ity density for the configuration of a single element x at
fixed y in the form
ρ1(x; y, β) = Z1(y, β)−1 exp [−βv(x; y)] ,
(2)
FIG. 1. Huxley-Simmons (HS) model of a single cross-bridge:
(a) mechanical representation a myosin head and (b) energy
landscape representing two chemical states. The conforma-
tion is characterized by the spin variable x which represents
two conformations of the head. The bistable element is linked
in series with a linear spring of stiffness κ0.
by the behavior of a single element. This justifies the ap-
proach of HS who could model the evolution of the first
moment of the underlying probability distribution by a
single reaction equation.
While we did not attempt in this paper to conduct a
systematic quantitative comparison of our statistical HS
model with experiment, we included at the end of the
paper a brief discussion of the relevance of our results for
skeletal muscles and for several other allosteric system
with mean-field coupling.
The paper is organized as follows. In Sec. II we study
the equilibrium properties of N bi stable elements con-
nected in parallel and loaded in a hard device. Sec. III
contains the analysis of the mechanical transients in this
system. The applicability of the original HS model for
the description of skeletal muscles is discussed in Sec. IV.
Various non-muscle applications are briefly reviewed in
Section V. In Sec. VI we summarize our results and iden-
tify some open problems.
II. EQUILIBRIUM
In this section we study the finite temperature equilib-
rium mechanical response of a folding-unfolding system
containing a finite number of elements.
A. Single HS element
The Huxley-Simmons paper [1] deals essentially with
a single folding element (representing a myosin cross-
bridge). The HS element can be modeled as an elastic
spring with stiffness κ0 (denoted by K in Ref. [1]) which
is connected in series with a bistable unit, see Fig. 1. The
two states represent the two conformations of the myosin
head and the variable x (denoted by −θ in Ref. [1]) takes
the values 0 (pre-power-stroke or unfolded conformation)
and −a (post-power-stroke or folded conformation). The
discrete "digital" nature of the conformational state in
κ0x(a)yκ0x0vHS(x)v0−a(b)By differentiating Eq. (4) with respect to y we obtain
the explicit representation of the equilibrium susceptibil-
ity
(cid:68)
[x − (cid:104)x(cid:105) (y, β)]2(cid:69)
χ(y, β) =
(cid:104)x(cid:105) (y, β) = β
∂
∂y
3
(5)
which is always positive, as expected in paramagnetic
systems. Given that the elastic element is linear, Eq. (5)
does not depend on the particular form of the energy
vHS(x). Thus, it also applies to models with more than
two discrete states [7] and even to models with continuous
energy landscape [2, 25].
Note that the susceptibility is proportional to the vari-
ance of x which in the HS model takes the form
[x − (cid:104)x(cid:105) (y, β)]2(cid:69)
(cid:68)
= (1/4){sech [β (y − y0) /2]}2 .
Both quantities will be used in what follows to assess the
intensity of fluctuations.
In the zero-temperature limit the variance of x is neg-
ligible at large absolute elongations. Instead, at y = y0,
the strength of fluctuations is independent of tempera-
ture and we obtain that χ = β/4, which is an analog
of the Curie law in paramagnetism [28]; see Figs. 2(c)
and 2(d). For other values of elongation y (cid:54)= y0, one can
define a characteristic temperature β = β∗χ(y) solving
the equation β∗χ(y − y0) tanh(cid:2)β∗χ(y − y0)/2(cid:3) = 1. At this
temperature fluctuations are maximized, see Fig. 2(d).
Below the characteristic temperature the system is es-
sentially "frozen" and therefore resistant to fluctuations.
Fluctuations are also irrelevant at large temperatures
where the system is maximally disordered.
b. Mechanical behavior. The free energy of a single
HS element in a hard device can be computed explicitly,
f (y, β) = − 1
β
log [Z1(y, β)] =
y2 + v0 +
1
2
(cid:26)
y − y0
(cid:21)(cid:27)
2
(y − y0)
(cid:20) β
2
ln
2 cosh
.
(6)
− 1
β
Its dependence on elongation is illustrated in Fig. 3(a).
We observe that for β ≤ 4 (large temperatures) the free
energy is convex while for β > 4 (small temperatures)
it is nonconvex. The emergence of a "pseudo critical"
temperature β = βc = 4 in a paramagnetic system is a
result of the presence of the quadratic energy associated
with the "applied field" y.
To study the mechanical manifestations of the implied
"criticality" we introduce the tension σ = y − x experi-
enced by the series linear spring. Due to the presence of
the quadratic term y2 in the energy, the conjugate vari-
able to elongation y is not the average "magnetization"
(cid:104)x(cid:105) but the average tension (cid:104)σ(cid:105) which is a linear function
of (cid:104)x(cid:105) independently of the form of the potential (1).
The convexity properties of the free energy can be
obtained through the study of the averaged tension-
elongation relation which corresponds to Eq. (16) in
FIG. 2. Average conformation and susceptibility of a single
HS element in thermal equilibrium. (a) and (b), Average con-
figuration as a function of the applied elongation at different
temperatures (a) and as a function of temperature at differ-
ent elongations (b); (c) and (d), susceptibility as a function
of elongation at different temperatures (c), and as function of
the temperature for selected values of y (d).
where
β =
κ0a2
kbT
is the nondimensionalized inverse temperature and kb is
the Boltzmann constant. The partition function for a
single element is then
(cid:20)
− β
2
(cid:21)
(cid:19)(cid:21)
(cid:18) y2
2
Z1(y, β) = exp
(y + 1)2
+ exp
−β
+ v0
. (3)
From (2) we can compute the average conformation (cid:104)x(cid:105) =
x={0,−1}
xρ1(x; y, β), obtaining,
(cid:104)x(cid:105) (y, β) = − 1
2
1 − tanh
(y − y0)
,
(4)
(cid:21)(cid:27)
(cid:26)
(cid:20)
(cid:20) β
2
(cid:80)
which is the analog of Eq. (15) in Ref. [1] (where the
corresponding variable was denoted by −n2). In param-
agnetic interpretation, (cid:104)x(cid:105) (y, β) is the "average magne-
tization" conjugate to the "applied magnetic field" y.
The dependence of (cid:104)x(cid:105) on the relative elongation y−y0
is illustrated in Fig. 2(a). In the zero-temperature limit
the system driven through y follows the global minimum
of the internal energy (1) and the population of the wells
changes discontinuously at y = y0 [26]. As the tempera-
ture increases, the transition smoothens and in the limit
β → 0 we have (cid:104)x(cid:105) = −1/2 independently of the elonga-
tion as illustrated in Fig. 2(b).
−1−0.500.51−1−0.50y−y0hxiβ2410∞05101520−1−0.50βhxiy−y0-0.250+0.25−1−0.500.51012y−y0χβ2410∞0510152000.511.5βχ00.250.51y−y0(a)(b)(c)(d)4
FIG. 3. Thermal equilibrium properties of the HS model in a hard device for different values of temperature. (a) Helmholtz
free energy; (b) tension-elongation relations; (c) stiffness. Parameters are β = 2 (dotted), β = 4 (solid), β = 10 (dashed) and
β → ∞ (dash-dotted). In the limit β → ∞, corresponding to zero temperature, the stiffness κ diverges at y = y0, see (c).
Ref. [1],
(cid:104)σ(cid:105) (y, β) =
∂f
∂y
= σ0 +y−y0− 1
2
tanh
(cid:21)
, (7)
(y − y0)
(cid:20) β
2
where σ0 = v0. The dependence of (cid:104)σ(cid:105) on the elongation
y − y0 is illustrated in Fig. 3(b) for different values of the
temperature.
We observe that while the relation (cid:104)x(cid:105) (y) at fixed tem-
perature is always monotone, as it is supposed to be in a
classical paramagnetic spin system, see Fig. 2, the depen-
dence of the tension (cid:104)σ(cid:105) on its conjugate variable y can
be nonmonotone, see Fig. 3(b). Behind this nonmono-
tonicity is the fact that the equilibrium stiffness
κ(y, β) = ∂ (cid:104)σ(cid:105) (y, β)/∂y = 1 − χ(y, β)
(cid:68)
[σ − (cid:104)σ(cid:105) (y, β)]2(cid:69)
= 1 − β
= 1 − (β/4){sech [β (y − y0) /2]}2
(8)
is a sign-indefinite sum of two terms.
Equation (8) is a representation of the standard [31, 32]
decomposition of an elastic susceptibility into a Cauchy-
Born part associated with affine deformation κCB = 1,
and a fluctuation part associated with nonaffine defor-
mation, here κF = (β/4){sech [β (y − y0) /2]}2. Interest-
ingly, in the HS model the fluctuation-related term in (8)
does not disappear in the zero-temperature limit, produc-
ing a singular δ-function-type contribution to the affine
response at y = y0. At this value of the elongation the
global minimum of the elastic energy is not unique and
fluctuations are formally present even in the zero tem-
perature (purely mechanical) model. This can be viewed
as a manifestation of a glassy behavior [33, 34].
At finite temperatures the fluctuation-related contri-
bution to the elastic modulus has a standard tempera-
ture dependence in pure phases y − y0 (cid:29) 1 (softening).
Instead, we observe a rubber-elasticity-type hardening
type behavior around y = y0, see Fig. 4. In this mixture
region the negative entropic elasticity starts to dominate
the positive enthalpic elasticity at β > βc.
The "critical" temperature βc = 4 is defined by the
condition that the tension-elongation relation develops
In this state κ = 1 − β/4,
zero stiffness at y = y0.
which can be again viewed as the analog of the Curie law
in magnetism. Negative stiffness, resulting from non-
additivity of the system, prevails at subcritical tempera-
tures; in this range a shortening of an element leads to a
tension increase which can be interpreted as a metama-
terial behavior [2, 35]. At supercritical temperatures the
stiffness becomes positive, reaching asymptotically the
value κ = 1.
It is remarkable that while fitting their experimental
data HS found exactly the critical value β = 4 (which cor-
responds to the choice 4/α = 8 nm in the units adopted in
Ref. [1]), concluding implicitly that the state of isomet-
ric contractions is only marginally stable. The advan-
tages of this state are clear from Fig. 4: Small variations
of temperature generate large changes in stiffness which
can vary from positive to negative values and such tem-
perature dependence is almost insensitive to the small
FIG. 4.
Equilibrium stiffness as function of the tempera-
ture at different levels of elongation. In the low temperature
regimes (large β), an increase of temperature induces soften-
ing while at high temperatures (low β) it induces hardening.
Close to the critical point βc, small changes in temperatures
have a large impact on the value of the stiffness which may
even change its sign.
−0.4−0.200.20.40.511.5y−y0fβ2410∞−110.511.5y−y+σ−σ+y−y0hσiσ0−11−101y−y+y−y0κ(a)(b)(c)02040−1−0.500.51βcsofteninghardeningβκy−y000.150.3t
FIG. 5. The temperature dependence of the parameter do-
main where the HS system exhibits negative stiffness.
In
(a) and (b) we show the horizontal (y+ and y−) and verti-
cal (σ+ and σ−) boundaries of the bistable domain; in (c)
the dashed line is a parametric plot of (y−(β), σ+(β)) and
(y+(β), σ−(β)) for ranging from β = 4 to β → ∞. Solid line:
Tension-elongation relation corresponding to β = 10 and the
associated bistable domain is represented by the gray area.
changes in the stretching around y0. The "criticality" in
HS system at y0, however, is subdued in the hard device
ensemble, similar to the behavior of a van-der-Waals gas
under controlled volume. The physical picture here is
differs from the case of a ferromagnetic system under ap-
plied magnetic field where interactions and cooperativity
play an important role and zero susceptibility signals the
presence of a real critical point with diverging fluctua-
tions.
To characterize the metamaterial behavior at temper-
atures below critical, we define an interval [y−, y+] where
the stiffness of the system is negative. The boundaries y−
and y+ correspond to the zeros of the second derivative
(cid:20)√
of the free energy. For β > 4 we have
√
β − 4
(cid:20)√
β − √
β − 4
√
β − 4
√
β − √
β − 4
y+(β) − y0 =
y−(β) − y0 = − 1
(cid:21)
(cid:21)
β +
β +
√
1
β
log
log
β
(cid:112)
(cid:112)
1
2
1 − 4β−1 − 1
β
1 − 4β−1 +
1
β
(cid:20)√
(cid:20)√
(cid:21)
(cid:21)
√
β − √
√
β − √
β +
β − 4
β − 4
β − 4
β − 4
In the zero-temperature limit this interval collapses to a
single point y = y0. The equilibrium tensions σ− and σ+
corresponding to y+ and y− are given by
√
σ+(β) = v0 +
β +
log
2
√
log
σ−(β) = v0 − 1
They become equal to σ+ = v0 + 1/2 and to σ− = v0 −
1/2, when β → ∞, see Ref. [26] for more detail. The
evolution of the domain of metamaterial behavior with
temperature is shown in Fig. 5.
.
We finish this
subsection with the observation
(cid:104)σ(cid:105) (y, β) = y − (cid:104)x(cid:105) (y, β), we have
that
since
5
(cid:104)[σ − (cid:104)σ(cid:105) (y, β)]2(cid:105) = (cid:104)[x − (cid:104)x(cid:105) (y, β)]2(cid:105), which shows that
the fluctuations of tension originate from the fluctuations
of the conformation.
In this sense the "noisy" macro-
scopic force-elongation relations can be used as an ex-
perimental window into the microscopic behavior of the
system.
c. Thermal behavior. While the experiments on
muscles have been traditionally focused on the mechani-
cal response [36–43], our study suggests that measuring
the thermal or calorimetric response of such systems may
be at least as informative, see some existing work along
these lines on muscles in Refs. [44–50]. The statistical
HS model has a considerable predictive power in this re-
spect. For instance, the entropy of the HS element can
be computed explicitly,
log [Z1(y, β)] + log [Z1(y, β)]
s(y, β) = −β
∂
∂β
(cid:26)
= log
2 cosh
(cid:20) β
2
(cid:21)(cid:27)
(y − y0)
(cid:20) β
2
− β
2
(y − y0) tanh
(y − y0)
(cid:21)
(9)
.
and we illustrate the behavior of the function s(y, β) in
Figs. 6(a) and 6(b). We see that the degree of disorder
is maximal in the state of isometric contractions, y = y0.
Note also that the entropy depends on a single normal-
ized coordinate β (y − y0) combining both control param-
eters, temperature and displacement.
A measure of the dependence of the entropy on tem-
FIG. 6.
Entropy [(a) and (b)] and specific heat [(c) and
(d)] in thermal equilibrium represented as function of elonga-
tion at different temperatures [(a) and (d)] and as function of
temperature for different elongations [(b) and (d)].
−1−0.500.510.60.811.21.4y−σ−y+σ+y−σ−y+σ+β=4β→∞β→∞β=10β=10y−y0hσiσ0−0.200.2y+y−y−y0σ−σ+051015200.811.2βhσiσ0(c)(a)(b)−2−101200.20.40.6y−y0sβ2410∞0510152000.20.40.6βs00.250.51−2−101200.20.4y−y0c0510152000.20.4βcy−y0(a)(b)(c)(d)6
FIG. 7.
from the state y = y0 at different temperatures.
Isothermal heat released induced by a displacement
FIG. 9. The dependence of the coefficient γ on elongation
(a) and temperature (b).
(cid:26) β
2
(cid:21)(cid:27)2
,
(y − y0)
(cid:20) β
2
perature is the specific heat [28]
c(y, β) = −β
∂
∂β
s(y, β) =
(y − y0) sech
which is represented as function of y − y0 and β in
Figs. 6(c) and 6(d). As is typical for paramagnetic
systems, the specific heat depends only on the com-
bination β(y − y0). Since at y = y0, the entropy is
temperature insensitive [s(y0) = log(2)], the specific
heat vanishes. Similarly, at large elongations, the sys-
tems becomes more and more ordered and tempera-
ture changes no longer affect the entropy. As a result,
the specific heat is maximized at a characteristic value
of the temperature β = β∗c which solves the equation
β∗c (y − y0) tanh [β∗c (y − y0)/2] = 2.
To study the heat release associated with the change
of length we can use our knowledge of the entropy vari-
ation with y. We introduce the heat release Q(y, β) =
−β−1∆s(y, β), where ∆s(y, β) = s(y, β)−s(yin, β), is the
entropy change from the initial state yin. The function Q,
which is illustrated in Fig. 7, can be potentially measured
by calorimetric techniques if the system is first driven
away from equilibrium adiabatically by a rapid length
change and then allowed to relax reaching the original
temperature.
Note that the expression for entropy (9) can be also
rewritten in the form s = β (cid:104)v(cid:105) − βf where (cid:104)v(cid:105) is the
FIG. 8. Average internal energy for β = 2 (dotted), β = 4
(solid), β = 10 (dashed) and in the athermal limit β → ∞
(dot-dashed).
average internal energy
(cid:104)v(cid:105) (y, β) = y2/2 + v0 − (y − y0)(cid:104)x(cid:105) (y, β).
In contrast to the equilibrium free energy, which de-
creases with temperature, see Fig. 3(a), the average in-
ternal energy increases with temperature, see Fig. 8 . In
the opposite zero temperature (athermal) limit (β → ∞)
both the average internal energy and the free energy tend
to the same limiting curve representing the global mini-
mum of the elastic energy which is a nonconvex function
of elongation energy [26]. Observe, however, that the av-
erage internal energy approaches the mechanical energy
"from above" while the free energy approaches the me-
chanical energy "from below".
Another interesting and potentially measurable quan-
tity is the entropic contribution to stress which also serves
as a measure of thermal expansion
γ = − ∂s
∂y
=
β2
4
(y − y0)
sech
(cid:26)
(cid:21)(cid:27)2
(y − y0)
.
(cid:20) β
2
The dependence of γ on elongation and temperature
is illustrated in Fig. 9. We observe that for y > y0
(respectivelyy < y0) the growth of temperature enhances
(respectively diminishes) the tension, see Fig. 3(b), and
the temperature sensitivity of tension is the highest at
a particular value of the temperature. In large shorten-
ing or stretching regimes and at y = y0, the mechanical
response is temperature insensitive.
d. Adiabatic response. The knowledge of the ther-
mal properties of the HS model allows one to address
the question of whether the isothermal approximation
is justified when applied to experiments involving fold-
ing or unfolding under fast loading. Below, we consider
an alternative hypothesis that the response is adiabatic,
which implies that in this problem the heat exchange is
the rate-limiting process. To remain within the equilib-
rium framework, we replace the task of computing the
actual adiabats by computing the isoentropes to which
we will be still referring as adiabats. We discuss the ap-
plicability of the adiabatic assumption for the description
of fast force recovery in muscles in Sec. IV.
As the entropy of the system depends solely on βy −
y0, see Eq. (9), the temperature varies along the adiabats
−2−101200.050.10.15β=4β=10β→∞y−y0Q−0.4−0.200.20.411.5y−y0hviβ2410∞−1−0.500.51−202y−y0γβ2410∞024681000.511.5βy−y000.250.51(a)(b)7
FIG. 10. Adiabatic response. (a) Evolution of temperature as function of the applied loading along adiabats for s = 0.2
(dotted), s = 0.4 (solid), and s = 0.6 (dashed). (b) Average conformation following adiabatic length changes from two different
thermal equilibrium initial conditions y = y0 (solid) and from y = y0 ± 0.1 (dashed) at βin = 10. The isothermal response is
represented by the dotted line. (c) Adiabatic tension-elongation relations. Dotted line, isotherm response for β = 10; dashed
line, adiabatic response with initial state at y − y0 = ±0.1 with βin = 10; solid line, adiabatic response with initial state at
y = y0 with βin = 10.
proportionally to the elongation, see Fig. 10(a). More
specifically, along an adiabat starting at y = yin with
temperature β = βin, we have βad = βin yin−y0
. Since,
y−y0
according to Eq. (4), the average configuration depends
only on β(y − y0), the variation of temperature along
adiabats must ensure that the average configuration (cid:104)x(cid:105)
is preserved. This is true for every value of y except y =
y0 where the adiabat experiences a discontinuity. Along
adiabats the average configuration evolves according to
sign(y − y0)
(cid:20) βinyin − y0
(cid:104)x(cid:105)ad (y, β) = − 1
+
tanh
(cid:21)
.
1
2
2
2
The adiabatic response of the microconfiguration of
the system to abrupt "length steps" is illustrated in
Fig. 10(b), where the initial temperature is always βin =
10. Observe that for the adiabat passing through the
point y = y0 the average configuration is frozen at
(cid:104)x(cid:105) = −1/2 (solid line); the behavior of a microconfigura-
tion along an isotherm passing through y = y0 drastically
differs (dotted line).
Since equilibrium tension along the adiabats depends
linearly on (cid:104)x(cid:105), the adiabatic stress response to shorten-
ing from y = y0 is quasilinear elastic, even though the
temperature is changing. More specifically, one can show
that outside the point y = y0 the adiabatic stiffness is
equal to the purely mechanical stiffness
κad =
∂2
∂y2
f (y, βad(y, s)) = κ0 ≥ κ,
where the function βad(y, s) describes temperature vari-
ation with elongation at a given entropy s.
Note that at y = y0, the inverse temperature β diverges
and even small adiabatic length change would lead to a
dramatic increase of temperature (β → 0). This means,
in particular, that reaching this state adiabatically brings
about infinite cooling. A Similar effect in a paramagnetic
spin system is known as "cooling by adiabatic demagne-
tization." In the HS system the applied field y − y0 can
be both positive and negative and, in this case, if y < y0,
a shortening would lead to a similar "adiabatic heating,"
which can be, in principle, measured in experiment, see
Section IV.
For the adiabats starting at other points y (cid:54)= y0 the
average configuration (cid:104)x(cid:105) is frozen at its initial value un-
til the loading reaches the point y = y0. At this point
the continuity of entropy requires that the configuration
changes discontinuously. Due to adiabatic cooling at
y − y0 the temperature goes to zero and the response be-
comes discontinuous (quasimechanical, see [26]). This is
in stark contrast with continuous evolution of the config-
uration along a typical isotherm also shown in Fig. 10(b)
(dotted line). One can say that, during adiabatic re-
sponse, the temperature-induced smoothing of the force-
elongation relation gets overridden by the anomalous
cooling around the point y = y0.
The adiabatic tension-elongation relations originating
from this behavior of the microconfiguration are piece-
wise linear with stiffness equal to 1 for y (cid:54)= y0, see
Fig. 10(c). The presence of a discontinuity at y = y0
signifies an extreme metamaterial-type behavior. Inter-
estingly, the adiabat originating exactly from the equilib-
rium state y = y0 can be confused with the purely elastic
isothermal force elongation relation. The associated tem-
perature variation, however, is non-negligible and should
be, in principle, measurable in experiments.
B. Bundle of HS elements
Consider now a finite number of HS elements attached
in parallel between two rigid backbones. In the skeletal
muscle context, such a bundle represents a minimal acto-
myosin complex which we refer to as an elementary half-
sarcomere, see Fig. 11.
The energy of the system with N elements can be writ-
−1−0.500.51051015200.20.40.6log(2)y−y0β−1−0.500.51−1−0.50isothermβ=10adiabatfromy=y0±0.1adiabatfromy=y0y−y0hxi−1−0.500.510.60.811.21.4isothermβ=10adiabatfromy=y0±0.1adiabatfromy=y0y−y0hσiσ0(a)(b)(c)8
which, in our case, plays the role of an order parame-
ter. The internal energy (per element) corresponding to
a given p can be written as
e(p, y) = p (y + 1)2 /2 + (1 − p)(cid:0)y2/2 + v0
(cid:1) .
(11)
(cid:18) N
(cid:19)
Due to permutational invariance, we can write the prob-
ability of a given state with N p elements in the folded
state in the form of the binomial law:
ρ(p; y, β) =
[ρ1(−1; y, β)]N p [ρ1(0; y, β)]N (1−p) ,
(12)
where ρ1 is given by (2) and ρ1(0; y, β) = 1−ρ1(−1; y, β).
Note that the distribution (12) can be also written as
N p
FIG. 11.
(a) Schematic representation of an acto-myosin
filaments organization in a superstructure of half-sarcomeres.
(b) A single half-sarcomere represented as a of cluster con-
taining N cross-linkers, see Fig. 1. The control parameter is
the total elongation y and the total tension generated by the
system is denoted by Σ.
ρ(p; y, β) = Z(y, β)−1 exp[−βN f (p; y, β)]
f (p; y, β) = e(y, p) − (1/β) s(p),
is the marginal free energy, e is the internal energy (11),
and s(p) = 1
sponding to a fixed value of p and finite N .
(cid:1) is the ideal entropy, all corre-
N log(cid:0) N
N p
(13)
ten as
e(x; y) =
1
N
N(cid:88)
(cid:20)
(1 + xi) v0 +
(cid:21)
,
(y − xi)2
1
2
where
i=1
where x = {x1, . . . , xN}. The individual bistable ele-
ments do not interact among themselves while they all
interact with the same external field y. The origin of this
mean-field type interaction is a hard device constraint
which is not affected by the microconfiguration of the sys-
tem. In the language of magnetism, we are dealing here
with a one-dimensional paramagnetic system.
In fact,
for such systems, the dimensionality is irrelevant and one
can expect the results obtained for the zero-dimensional
model to remain valid for the case of N elements.
e.
Thermal equilibrium.
In thermal equilibrium,
the probability density for a micro-state x reads
ρ(x; y, β) = Z(y, β)−1 exp [−β e(x; y)] ,
(10)
where the partition function is
Z(y, β) =
exp [−βN e(x; y)] .
(cid:88)
x∈{0,−1}N
ρ(x; y, β) = (cid:81)N
Due to the additivity of the energy we obtain Z(y, β) =
[Z1(y, β)]N , where Z1 is given by (3).
Therefore
i=1 ρ1(xi; y, β) which shows that the el-
ements are independent.
The total free energy can be written as F (y, β) =
N f (y, β), where the expression for the free energy of a
single HS element f is given by (6); this formula is anal-
ogous to the corresponding result for paramagnetic Ising
model and other mean-field-type systems, e.g., Ref. [16].
Similarly, other extensive equilibrium variables are also
additive and it will be convenient to normalize them by
N .
To shed light on the internal microconfiguration of the
system we introduce the fraction of HS elements in the
folded conformation
N(cid:88)
i=1
xi,
p = − 1
N
FIG. 12. Nonequilibrium free energy landscape and the cor-
responding tension-elongation relations in a hard device for
N = 2 [(a) and (b)] and in the limit N → ∞ [(c) and (d)].
Dotted lines, free-energy levels, and tension corresponding to
different values of p; thick line, response corresponding to the
global minimum of the nonequilibrium free energy; thin lines,
response in thermal equilibrium; gray areas, domain of the
metastable states in the thermodynamic limit. In (a) and (c),
the inserts show the marginal free energy f as function of p
for y − y0 = −0.4, 0, 0.4. The plots are obtained with β = 6,
which explains the presence of negative stiffness.
half-sarcomere(a)ΣΣrigidbackboneuHSκNy(b)0.511.52fN=2N→∞−0.4−0.200.20.4012y−y0⟨σ⟩σ0−0.4−0.200.20.4y−y001pf01pf01pf01pf∞01pf∞01pf∞(a)(b)(c)(d)9
FIG. 13. Hill-type energy landscapes for N = 1 (a) and N = 4 (b). In (c) we show the equilibrium free-energy profile f = F/N
(solid line), which is independent of N together with the metastable states for N = 4 (dotted lines). Here v0 = 1/2.
To illustrate Eq. (13), consider the simplest case N = 2
when the marginal free energy can take only three values,
f (0; y, β) = e(0; y) =
1
2
y2 + v0
f (1/2; y, β) =
1
4
(y + 1)2 +
1
4
y2 +
1
2
v0 − log(2)
2β
f (1; y, β) = e(1; y) =
1
2
(y + 1)2,
which are shown in Fig. 12 together with the correspond-
ing tension-elongation relations. Observe that the global
minimum response at finite temperature is characterized
by a series of jumps reflecting successive conformational
changes in individual elements. Between the jumps, the
stiffness is positive, which shows that each metastable
state has a finite basin of stability even though the over-
all (global) stiffness is negative.
The changes in the marginal free-energy profiles with
increasing N are illustrated in Fig. 13. At N = 1, we ob-
tain the representation of the energy landscape due to T.
L. Hill [36]. In this case, the marginal free energy f and
the internal energy v are identical (no entropic contri-
bution). For finite N we obtain N + 1 metastable states
corresponding to different values of p with the global min-
imum represented by a (nonconvex) lower envelope.
While the lower envelope of the marginal free energy
f (y, β) = minp
f (p, y, β) is a piece wise smooth function
of y with a number of singular points depending on N , the
equilibrium free energy f (y, β) = F (y, β)/N is a smooth
function laying strictly below: f (y, β) ≤ f (y, β). The N
independence of f (y, β)- see Eq. (6)-shows that for the
HS system the equilibrium response is size independent.
However, in real experiments for systems with small N
conducted at finite deformation rates one can expect to
see the steps on the force elongation curves associated
with the singularities of f (y), see Sec. III B.
The average value of the parameter p (which is analo-
gous to the variable n2 in Ref. [1]) can be found from
p ρ(p; y, β) = −(cid:104)x(cid:105)(y, β) = ρ1(−1; y, β)
(cid:104)p(cid:105) (y, β) =
(14)
(cid:88)
This quantity plays the role of the average magnetiza-
tion per spin and does not depend on N ; however, the
corresponding variance decreases as 1/N ,
(cid:104)[p − (cid:104)p(cid:105) (y, β)]2(cid:105) = (1/N )(cid:104)[x − (cid:104)x(cid:105) (y, β)]2(cid:105).
Our Eq. (14) also shows that the whole distribution (12)
can be recovered if the parameter N is fixed and (cid:104)p(cid:105) is
known as a function of y and β. In particular, we can
compute the variance
(cid:104)[p − (cid:104)p(cid:105) (y, β)]2(cid:105) = (cid:104)p(cid:105) (1 − (cid:104)p(cid:105)),
(15)
which gives after substitution
(cid:68)
[p − (cid:104)p(cid:105) (y, β)]2(cid:69)1/2
=
√
1
N
2
sech
(cid:21)
(y − y0)
(cid:20) β
2
.
(16)
By differentiating Eq. (14) with respect to y we can also
obtain the equilibrium susceptibility
X(y, β) = − ∂
∂y
(cid:104)p(cid:105) = βN(cid:104)(p − (cid:104)p(cid:105))2(cid:105) = χ(y, β),
where χ is the susceptibility of a single HS element,
see Eq. (5). We can similarly rewrite all other equilib-
rium characteristics of the system in terms of −(cid:104)p(cid:105) and
N(cid:104)[p − (cid:104)p(cid:105) (y, β)]2(cid:105).
In the limit N → ∞ the expression for the marginal
free energy can be written explicitly
f∞(p; y, β) = e(p; y) − (1/β) s∞(p),
(17)
where s∞(p) = − [p log(p) + (1 − p) log(1 − p)] , is the
ideal mixing entropy reflecting the absence of correla-
tions between the units. The function f∞(p) is always
convex since
f∞(p; y, β) =(cid:2)β p(1 − p)(cid:3)−1 > 0,
∂2
∂p2
which signifies the lack of synchronization: In a similar
ferromagnetic system the marginal free energy would be
nonconvex. As N → ∞ the domain of the phase space
occupied by the metastable states becomes compact, see
prepost−2−10100.20.40.60.81y−y0vN=1p=0p=1/4p=1/2p=3/4p=1−2−10100.20.40.60.81y−y0fN=4−2−10100.20.40.60.81y−y0f∀N(a)(b)(c)the gray area in Figs. 12(c) and 12(d) not shown explicitly
in Fig. 13(c).
where the superscript d indicates that the normalization
has been dropped.
10
At large N , the summation over the set of discrete
values of p (see Eq. (14)) can be approximated by an
integration over the interval [0, 1]. The integrals can be,
in turn, computed by using the Laplace method. Then,
for the equilibrium free energy, we can write f (y, β) =
f∞(p∗(y, β); β), where f and f∞ are given by (6) and
(17), respectively. Here p∗(y, β) is a minimizer of f∞,
which is a solution of the transcendental equation,
p∗/(1 − p∗) = exp(cid:2)−β(y − y0)(cid:3).
It is easy to check that p∗(y, β) = (cid:104)p(cid:105) (y, β) where
(cid:104)p(cid:105) (y, β) is given by Eq. (14) . The resulting free-energy
profile is shown in both Fig. 12(c) and Fig. 13(c).
f. Mechanical behavior. For a given configuration x,
the tension in the system can be written as
(cid:104)
(cid:88)
(cid:105)
Σ(x, y) = N
y − (1/N )
xi
= N (y + p) .
The average tension, conjugate to the control parameter
y, is then
(cid:104)Σ(cid:105) (y, β) = N [y + (cid:104)p(cid:105) (y, β)] = N (cid:104)σ(cid:105) (y, β),
where (cid:104)σ(cid:105) is the average tension of a single element, see
Eq. (7). The variance of the total tension can be written
as
(cid:68)
[σ − (cid:104)σ(cid:105) (y, β)]2(cid:69)
.
(cid:68)
[Σ − (cid:104)Σ(cid:105) (y, β)]2(cid:69)
(cid:68)
[Σ − (cid:104)Σ(cid:105) (y, β)]2(cid:69)1/2
= N
The relative fluctuations,
(cid:26)
(cid:104)Σ(y, β)(cid:105)
(cid:20) β
2
(y − y0)
=
(cid:21)
− 1
2
(cid:20) β
2
sinh
(18)
(cid:21)(cid:27)−1
(y − y0)
√
1
N
2
(y +
) cosh
1
2
decay as 1/N 1/2, which is a sign that the measured force
in this model is an extensive quantity. The formula (18)
can be used to estimate the number of elements N from
the knowledge of the fluctuations of the force.
If we denote by K the total stiffness of the system,
then we can write
K(y, β) = N κ(y, β) = N − β
(cid:68)
[Σ − (cid:104)Σ(cid:105) (y, β)]2(cid:69)
(19)
Note, first, that the fluctuation-related contribution to
stiffness is an order of one effect in terms of the number
of elements N , which means that the effect of fluctu-
ations does not disappear in the thermodynamic limit.
Also, since the stiffness in the HS model is an extensive
property, the analysis of the temperature and elongation
dependence of κ(y, β) presented in Sec. II A remains valid
here as well.
The fact that the stiffness has an nonthermal, purely
mechanical part, that can be potentially extracted from
structural measurements, and an equally important and
even dominating fluctuation-related component, that can
be measured independently, has been largely overlooked
in the literature on systems with nonconvex internal de-
grees of freedom because in classical materials, which can
be thought to be composed of almost linear springs, the
fluctuational effect on stiffness is usually small. Here we
see that in the presence of internal "snap-springs" this ef-
fect can be considerable. For instance, the difference be-
tween the smaller quasistatic stiffness of myosin II [51, 52]
and the larger instantaneous stiffness-believed to be
largely unaffected by fluctuations [53] -may be linked
to the importance of the second term in our Eq. (20).
On the other hand, if N is known and the variance of
the total force can be measured, then one can recover the
stiffness of a single element κ0. Conversely, knowing κ0
and measuring fluctuations of the force one can estimate
N from (20). We emphasize again that the relation (20)
is independent of the detailed structure of the energy
landscape (1) and can therefore be used in the presence
of multiple power-stroke-type energy wells.
g. Thermal behavior.
Since the entropy of the fi-
nite size bundle of HS elements is extensive S(y, β) =
N s(y, β), the analysis of the adiabatic response for a sin-
gle HS element presented in Sec. II A remains valid for
the bundle of N HS elements.
III. KINETICS
In this section we study kinetics of the HS system and
build links between the stochastic dynamics of a single
HS element, the evolution of the bundle of HS elements
connected in parallel, and a conventional chemomechan-
ical modeling of such systems in terms of deterministic
chemical reactions. Following the original HS model, we
assume for simplicity that during loading the tempera-
ture is kept constant.
where κ is defined by Eq. (8). As in the case of a single HS
element, the total stiffness decomposes into an elastic (or
enthalpic) contribution dominating at y − y0 (cid:29) 1 and
a term containing entropic contribution which dominates
around y = y0.
(cid:28)(cid:104)
In dimensional form (19) becomes
(cid:34)
K d = N κ0 − 1
kbT
1 − κ0 a2
4kbT
Σd − (cid:104)Σ(cid:105)d(cid:105)2(cid:29)
(cid:20) κ a
(cid:26)
= N κ0
2kbT
sech
A. Single HS element
(cid:1)(cid:21)(cid:27)2(cid:35)
,
(20)
In the paper of Huxley and Simmons [1] the relaxation
of the system to equilibrium was modeled as a determin-
istic chemical reaction of the first order.
In their de-
scription HS followed the average population of elements
(cid:0)yd − yd
0
FIG. 14. One-dimensional Markov chain description for a
single HS element (a) and for a system with N elements (b).
in the two conformational states without attempting to
trace the dynamics of individual flips experienced by the
spin variables x.
To simulate stochastic dynamics of a single HS element
we need to know the probabilities of the forward and
reverse flips
P(cid:2) xt+dt = −1(cid:12)(cid:12) xt = 0(cid:3) = k+(y, β)dt
P(cid:2) xt+dt = 0(cid:12)(cid:12) xt = −1(cid:3) = k−(y, β)dt.
(21)
Here k+(y, β) [respectively k−(y, β)] is the transition rate
for the jump from the unfolded state (respectively, folded
state) to the folded state (respectively, unfolded state),
see Fig. 14(a). We assume that the total elongation y
and the inverse temperature β are the controlling pa-
rameters which may vary at a time scale much larger that
the characteristic time of the individual conformational
transitions. As the transition probabilities (21) depend
only on the current state of the system, the dynamics is
described by a discrete Markov chain [30, 54].
To compute the transition rates k±(y, β) we need to
know the structure of the actual energy landscape sepa-
rating the two conformational states. In their paper [1],
HS simply assumed that the hypothetical barrier sepa-
11
rating the two wells of the potential vHS is flat and is
characterized by the energy level E1, see Fig. 15. By
taking the energy of the elastic spring into account, we
can then write the transition rates in the form
k+(y, β) = k exp [−β [E0 + max{y + 1/2, 0}]] ,
k−(y, β) = k exp [−β [E1 + max{−y − 1/2, 0}]] .
where E1 = E0 + v0 and the common pre-factor k defines
the characteristic time scale for a single well system. If
y > −1/2, then only the energy barrier from the unfolded
to the folded state depends on y, see Fig. 15(a). In this
case, the rates can be written in the form
k+(y, β) = k− exp [−β (y − y0)]
k−(y, β) = k exp [−β E1] = const
(22a)
(22b)
and the timescale of the jump process is τ = 1/k− =
k−1 exp [β E1] . If y < −1/2, see Fig. 15(b), then we ob-
tain instead
k+(y, β) = k− exp [β v0] = const.
k−(y, β) = k− exp [β (y − y0 + v0)] .
We see in Fig. 15 that,
in response to shortening,
the overall transformation rate k = k+ + k− first in-
creases exponentially as the forward barrier is lowered
(while the reverse barrier remains constant) and then
decreases as the reverse barrier is elevated (while the
forward barrier remains constant).
In addition, we see
that for large stretching k ≈ k− and for large shortening
k ≈ k− exp[β v0].
Note that HS considered only the case y > −1/2, see
Fig. 15(a). To see why the value y = −1/2 is special in
the muscle context, we recall that in this case the un-
loaded system is symmetric, (cid:104)p(cid:105) = 1/2. Then the tension
during the purely elastic phase of the fast force recovery
(when (cid:104)p(cid:105) remains constant) is σ = y + 1/2, which be-
comes negative exactly at y = −1/2. In experiments on
muscles, the relaxation rates have been measured only
for y > −1/2 because below this threshold muscle fibers
are usually subjected to buckling. Our analysis suggests
that near the regimes with y = −1/2 the step size depen-
dence of the rate of fast force recovery may deviate from
exponential.
Schematic representation of the energy barrier in
FIG. 15.
the HS bistable potential. The energy barriers corresponding
to the transition rates in the absence of elastic contribution
are denoted E1 and E2. We define the characteristic timescale
by τ = exp[βE1]. (a) Energy landscape for y > −1/2 which
is the case considered in Ref. [1]; (b) energy landscape for y <
−1/2 not considered by HS. (c), Relaxation rate as function
of the total elongation y for β = 1 (dashed line) and β = 2
(solid line).
The results of numerical simulations of stochastic hop-
ping for a single HS element subjected to a quasistatic
stretching are shown in Fig. 16. For y < y0 the spin vari-
able spends most of the time in the folded conformation
(x = −1). When the loading device approaches the point
y = y0, the flips between the wells become more frequent
before finally the system stabilizes again in the unfolded
configuration (x = 0).
The stochastic dynamics shown in Fig. 16 can also
be seen through the prism of the deterministic evolution
of a single-particle probability distribution ρ1(t). For a
generic test function q we can write
d(cid:104)q(x)(cid:105) = q(−1) [ρ1(−1, t + dt) − ρ1(−1, t)]
+ q(0) [ρ1(0, t + dt) − ρ1(0, t)] ,
(a)−10k+k−(b)p−1Npp+1Nφ+(p−1N)φ+(p)φ−(p+1N)φ−(p)x0v0E0E1−1(a)y>−1/2ux0v0E0E1−1(b)y<−1/2u−2−101246β=1β=2ykτ(c)12
between the time t and the time t + dt, then the func-
tion p(t) is a one dimensional random walk (see Fig. 14),
governed by the jump probabilities
(cid:40)P(cid:2)pt+dt = pt + 1/N(cid:3) = φ+(pt; y, β)dt
P(cid:2)pt+dt = pt − 1/N(cid:3) = φ−(pt; y, β)dt.
(24a)
(24b)
where φ+(p; y, β) = N (1 − p) k+(y, β) and φ−(p; y, β) =
N p k−(y, β). Following a similar procedure as the one
leading to Eq. (23), we obtain the master equation for
the probability distribution ρ(p, t)
ρ(p, t; y, β) = φ+ (1 − p + 1/N ; y) ρ (p − 1/N, t; y, β)
∂
∂t
+ φ− (p + 1/N ; y) ρ (p + 1/N, t; y, β)
− [φ+(1 − p; y) + φ−(p; y)] ρ (p, t; y, β) ,
(25)
which can be solved numerically since we know the tridi-
agonal transfer matrix of the process at each time step.
It is clear, however, that since the transition probabilities
(21) depend only on the control parameter y, the trajec-
tories of individual elements are independent. Hence, at
a given y each macro-configuration can be viewed as a
realization of N Bernoulli processes with the probability
of success ρ1(t) solving Eq. (23). Therefore the probabil-
ity density ρ(p, t) = P(pt = p) is a binomial distribution
with parameters N and ρ1(t):
(cid:18) N
(cid:19)
N p
FIG. 16.
Hopping response to a ramp stretch from y =
y0 − 1 to y = y0 + 1 (single trajectory with jumps hardly
distinguishable around y = y0 ). The time step is ∆t = 10−3τ
and the time is measured in the units of τ . The average
trajectory is shown by the gray line. Here β = 4.
then, using (21), we obtain
d(cid:104)q(x)(cid:105) = q(−1){k+[1 − ρ1(−1, t)] − k−ρ1(−1, t)} dt
+ q(0){k−[1 − ρ1(0, t)] − k+(ρ1(0, t))} dt.
In the limit dt → 0, we obtain exactly the HS kinetic
equation
∂
∂t
ρ1(t) = k+(y) [1 − ρ1(t)] − k−(y)ρ1(t).
(23)
ρ(p, t) =
[ρ1(t)]N p [1 − ρ1(t)]N−N p .
(26)
Its general solution can be written as
ρ1(t) = ρ1(0) exp [−A(t)] +
(cid:90) t
0
where A(t) =(cid:82) t
k(t(cid:48)) exp[A(t(cid:48)) − A(t)]ρ∞1 (y(t))dt(cid:48),
0
k(t(cid:48))dt(cid:48) and ρ∞1 = k+/k is the stationary
distribution (2). Since (cid:104)x(cid:105) = ρ1, this equation describes
the time dependence of the average configuration shown
in Fig. 16 by the gray line.
The comparison of individual stochastic trajectories
with the evolution of averages shows that the information
about individual flips, potentially measurable in single
molecule experiments, gets lost in the chemomechanical
description. In particular, near the point y = y0, fluc-
tuations play a dominant role in the stochastic descrip-
tion, as is suggested by our equilibrium theory, while from
the chemomechanical perspective this particular state is
completely indistinguishable from the other equilibrium
states.
B. Bundle of HS elements
The isothermal discrete dynamics of a system with N
elements can be described in terms of the macroscopic pa-
rameter p ∈ {0, 1/N, . . . , 1}. If only one transition occurs
One can verify that that (26) solves (25) and since
(cid:104)p(cid:105) (t) = ρ1(t), Eq. (23) can be viewed as the analog of
Eq. (9) in Ref. [1]. We have then shown that the dynam-
ics of the entire distribution is enslaved to the dynamics
of the order parameter (cid:104)p(cid:105) (t) captured by the original HS
model. It is also clear that in the long time limit the dis-
tribution (26) converges to the Boltzmann distribution
(10).
h. Quasistatic loading. To illustrate the fact that
our dynamical model is fully compatible with the equi-
librium behavior studied in Sec. II B, we now consider
the quasistatic driving of a cluster of N HS elements,
see Fig. 17. The behavior of the individual trajecto-
ries generated by the stochastic random walk Eq. (24)
is shown for N = 10 (light gray) and N = 100 (dark
gray). The system is subjected to continuous stretching
from y = y0 − 1 to y = y0 + 1 over the time interval
[0, 103 τ ] with the temperature remaining constant; this
loading protocol mimics the unzipping tests for biolog-
ical macromolecules [12, 15, 18, 55]. The results were
obtained using the same numerical procedures as in the
case of a single element.
The stochastic evolution of the order parameter p and
of the corresponding tension are illustrated in Fig. 17.
Together with single trajectories, we show the evolution
of the average (solid black line) obtained from Eq. (23)
and the corresponding equilibrium response curves (open
−10x−101y−y005001000012t/τσσ013
FIG. 18. Evolution of the normalized variance N(cid:104)(p − (cid:104)p(cid:105))2(cid:105)
and the heat release as function of the elongation during a
quasistatic stretching between t = 0 and t = 103τ . For each
temperature we show the analytic computations from (16)
and (9) (open circles), the results of the stochastic model (24)
corresponding to 104 independent realizations with N = 10
(gray lines) and the solution based on the solution of the
kinetic equations (23) and (26) (thin lines).
that stochastic simulations are fully compatible with the
predictions of equilibrium theory, in particular, we see
once again that the normalized variance reaches a max-
imum at y = y0 becoming independent of the tempera-
ture.
i. Fast loading.
In addition to averages, captured al-
ready by the chemomechanical kinetic equation (23), the
master equation (25) allows one to follow the evolution
of higher order moments. To illustrate this point, we
now show how the HS system responds to abrupt per-
turbationss which is exactly the type of mechanical test
conducted in Ref. [1], see Fig. 19. The system is first
maintained in equilibrium at y = y0 = 0.5 before an in-
stantaneous length change (to y = 0.) is applied. This
protocol is repeated for systems with N = 10 (dotted
lines) and N = 100 (solid line). Again, individual real-
Stochastic simulation of a quick force recovery
FIG. 19.
in response to a step of y − y0 = −0.5.
(a) Tension per
cross-linker; (b) heat released. Dotted lines: Single stochastic
trajectory for a system with N = 10; gray line single trajec-
tory with N = 100; solid line average over 1000 trajectories
with N = 10; squares, response obtained using the HS kinetic
equation (23). Inserts in (a), gray (respectively, solid) line,
fluctuations obtained using 1000 realizations for N = 100 (re-
spectively, N = 10). Here β = 4 and y0 = 0.5.
FIG. 17. Response of the HS model to a ramp loading from
y0 − 1 to y0 + 1 achieved in 1000 τ , for β = 4 [(a) and (c)] and
β = 10 [(b) and (d)]. Individual stochastic trajectories are
shown for N = 10 (light gray) and N = 100 (dark gray). [(a)
and (b)] Evolution of the order parameter p. The inserts show
two samples of a single trajectory around the point where
p = 1/2.
[(c) and (d)] Tension-elongation relations obtained
from σ = y + p. The inserts show the marginal distribution
ρ at the two different times indicted by the vertical bars.
Solid lines represent the thermal equilibrium averages given
by Eq. (14) and Eq. (7) and open symbols show the solutions
of the HS kinetic equation (23) and the distribution (26).
circles). The inserts in Figs. 17(a) and 17(b) show sam-
ples of the trajectories for single elements computed from
(21).
Observe that individual trajectories reveal at finite
N a succession of jumps describing individual folding-
unfolding events as is suggested by the analysis of the
marginally equilibrated system. As the number of ele-
ment increases the fluctuations of p decrease in accor-
dance with our Eq. (16), and a single realization tra-
jectory (dark gray) gets close to the average trajectory
(black).
In the inserts in Figs. 17(c) and 17(d) we show the
probability density ρ obtained from (26) at different
times (solid line) together with the equilibrium density
(open circles). As expected, the distribution does not
depend on temperature at y = y0 while becoming pro-
gressively more localized away from this point.
To make the stochastic fluctuations more visible, we
compare in Fig. 18 the variance of the order parameter
p obtained from the stochastic model (24) (gray lines)
with the results of the analytic computations based on
the kinetic equation (26) (thin lines) and the equilibrium
model, see Eq. (16) (open circles). The system contains
N = 10 elements and each stochastic trajectory corre-
sponds to 104 realizations of our random walk. We see
0250500750100000.51t/τpβ=402505007501000t/τβ=10−1−0.500.510.60.811.21.4y−y0σσ0−1−0.500.51y−y0300400−10t/τxi600700−10t/τxi400500−10t/τxi500600−10t/τxi01pρ01pρ01pρ01pρ(a)(b)(c)(d)−1−0.500.5100.10.20.25β=2β=10y−y0Nh[p−hpi]2i−1−0.500.5100.050.1β=2β=10y−y0Q(a)(b)−10120.40.60.811.2t/τΣN−101200.050.10.15t/τQ020.10.251Nh(Σ−hΣi)2i(a)(b)14
FIG. 20. Time evolution (snapshots) of the probability den-
sity ρ(p, t; y, β) showing gradual equilibration of the system
subjected to an abrupt shortening. Lines: histograms ob-
tained from 104 trajectories; symbols distributions recovered
from the first order kinetic equations (23) and (26). Parame-
ters are the same as described in the caption to Fig. 19
izations may strongly depart, especially at low N , from
the average behavior described by the HS reaction equa-
tion (23) (symbols).
These fluctuations can also be seen from the dynam-
ics of the density ρ which is reconstructed from a large
number of sample trajectories/experiments in Fig. 20.
We observe that for a system with a small number of
elements [see Fig. 20(a), N = 10) the probability distri-
bution remains broad even after the recovery while, for a
system with large N [see Fig. 20(b), N = 10), the distri-
bution is sharply peaked throughout the process. Again,
we find a perfect agreement between the distribution ob-
tained from the Monte-Carlo simulations (lines) and the
one recovered from the knowledge of the averaged behav-
ior given by the HS kinetic equation parametrizing the
binomial distribution (26) (symbols).
IV. SKELETAL MUSCLES
The development of the HS model was originally mo-
tivated by the mechanical experiments involving rapid
shortening of skeletal muscles with the goal of distin-
guishing passive from active contributions to tension re-
covery [1, 56–59]. It was shown that the first phase of
the response to a quasi-instantaneous shortening imposed
on a maximally activated (tetanized) single muscle fiber
represents a purely elastic force drop. During the sec-
ond phase, the tension recovers to a level which depends
nonlinearly on the amplitude of the shortening. This fast
force recovery, lasting about 1 ms, precedes a consider-
ably slower phase at the end of which the tension fully
returns to its original value. The latter, taking place on
a 100-ms time scale, is usually interpreted as an active
process driven by ATP hydrolysis [7, 60].
j. Biochemistry vs mechanics.
In their classical
1971 paper HS conjectured that the force recovery at the
ms time scale must be attributed to a rapid folding in
an assembly of attached cross-bridges linking actin and
myosin filaments. The idea of bistability in the structure
FIG. 21. Biochemical vs purely mechanistic description of
the power stroke in skeletal muscles: (a) The Lymn-Taylor
four-state cycle, LT(71) and (b) the Huxley-Simmons two-
state cycle, HS (71).
of myosin heads, giving rise to the concept of a power
stroke, has been later fully supported by crystallographic
studies [61, 62].
While the scenario proposed by HS is in agreement
with the fact that the power stroke is the fastest step
in the Lymn-Taylor (LT) enzymatic cycle [63, 64], there
is a subtle formal disagreement with the existing bio-
chemical picture, see Fig. 21. Thus, HS assumed that
the mechanism of the fast force recovery is fully passive
and can be reduced to a mechanically induced confor-
mational change.
In contrast, the LT cycle for acto-
myosin complexes is based on the assumption that the
power stroke can be reversed only actively through the
completion of the bio chemical pathway including ADP
(adenosine diphosphate) release, myosin unbinding, bind-
ing of uncleaved ATP (adenosine triphosphate), splitting
of ATP into ADP and Pi, and then rebinding of myosin
to actin [6, 63].
In other words, while HS postulated that thermal fluc-
tuations experienced by the attached myosin heads can
be biased by external loading and that the power stroke
can be reversed by mechanical means, most of the bio-
chemical literature is based on the assumption that the
power-stroke recocking cannot be accomplished without
the presence of ATP. In particular, physiological fluctu-
ations in muscle response are mostly addressed in the
context of active behavior [65–71].
Some authors, however, follow the HS mechanistic ap-
proach in assuming that the power-stroke-related leg of
the LT cycle can be decoupled from the rest of the bio-
chemical pathway; see, for instance, Refs [49, 72]. Below
we adopt this perspective, which implies that at a 1-ms
time scale the mechanism dominating muscle response
is a purely mechanical folding-unfolding. We then col-
lect specific predictions, generated by our augmented HS
model, and use them as a guidance in designing new ex-
periments aimed, in particular, at verifying the correct-
ness of the underlying purely mechanistic model.
k. Thermal effects. The knowledge of the free en-
ergy of the HS system allows one to assess not only me-
chanical but also thermal manifestations of the fast force
recovery. The latter have been measured in experiments
employing calorimetric techniques [49, 50, 73], however,
a thermomechanical interpretation of these experiments
in the HS framework is still an open question.
00.5100.20.4t=0t=0.1τt=τpρN=1000.5100.050.1t=0t=0.1τt=τpN=100AM(AMADPPi)AM(AMADP)M(MATP)M(MADPPi)(a)LT(71)AMAM(b)HS(71)For instance, studies of the heat exchange following
the application of a fast length drop showed an increase
of temperature at the time scale of the purely elastic
response followed by a slower cooling during the force
recovery up to a level which is higher than the baseline
preceding the step. While, the temperature decay was
linked to the equilibrium heat effect of the conformational
change, which was assumed to be negative [73], the HS
model predicts a positive heat effect because a "mixed"
state with high entropy is transformed into a "pure" state
with low entropy.
More specifically, the HS-type interpretation of the
temperature measurements during the fast force recov-
ery would be as follows:
(i) The rapid increase of temperature recorded dur-
ing the applied length step is a reflection of an adiabatic
temperature increase. To justify this claim, we mention
that several experimental studies of muscles, involving
temperature changes due to rapid switching between, so-
lutions showed that the time scale of temperature equi-
libration within a typical muscle fiber is of the order of
10 ms [74, 75], which is 10 times slower than the duration
of the fast force recovery process.
(ii) The subsequent temperature decay is an outcome
of the cooling due to heat conduction and the heat release
due to the conformational change. The fact that the
temperature at the end of the recovery is higher than
the baseline temperature is a signature of the exothermic
nature of the folding process (of the working-stroke) and
the inefficiency of the heat removal mechanism at this
time scale.
Several groups have also addressed the influence of
temperature on the force generation either by perform-
ing mechanical experiments in different solutions [45, 46]
or by applying rapid temperature changes to tetanized
muscle fibers [47, 48, 76]. In both cases the experiments
show that the isometric tension increases with tempera-
ture while the conformational state of the cross-bridges
becomes more homogeneous. Such a response in the case
of fast adiabatic changes can be explained by the fact
that, in order to maintain the value of the entropy at
higher temperature, the HS system must evolve towards
a more ordered configuration. This effect, however, is not
captured by the HS model where temperature does not
affect the value of isometric tension at y = y0, see our
Fig. 3. To describe quantitatively the temperature de-
pendence of the tension-elongation curves, we may aug-
ment the HS model by assuming phenomenologically that
the energy bias v0 is a function of temperature [46].
While the value of the isometric tension may depend
significantly on temperature, experiments show that the
slope of the tension-elongation curve is only weakly tem-
perature sensitive [49].
In general, the HS model pre-
dicts considerable dependence of the shape of the tension-
elongation relation on temperature, including the possi-
bility of negative stiffness which is not observed in exper-
iment. However, in the range of temperatures considered
in experiment (no more than 20°C) this effect is weak.
15
For instance, if we assume that the mechanical properties
of the cross-bridges are not affected by temperature, as is
observed in experiments [45], and take β = 4 at T = 4°C,
we obtain β ∼ 3.8 at T = 24°C. In the HS model, the
sensitivity of the equilibrium tension-elongation curves
to such variations of β is negligible.
Among other interesting thermomechanical effects, in-
voked by the HS model, we mention the "infinite" cooling
in the process of reaching the state of isometric contrac-
tions. To observe this effect the muscle must be first
equilibrated after shortening (at the T2 state) and then
stretched back to the T0 state.
l. Fluctuations. While in myofibrils half-sarcomeres
fluctuations can be expected to average out, at the
scale of an elementary acto-myosin complex (our half-
sarcomere, see Fig. 11), where N ∼ 102, fluctuations may
interfere with experiments. For instance, as we have seen
in our Fig. 17 and Fig. 19, the abrupt transitions asso-
ciated with conformational changes in individual cross-
bridges may produce measurable steps in the response
curves.
Another way of assessing the role of fluctuations is
through the measurement of equilibrium susceptibilities.
Thus, the effective stiffness of the HS bundle can be rep-
resented as a sum of an enthalpic term describing zero-
temperature elasticity and an entropic term that can be
evaluated from the measurements of tension fluctuations.
More specifically, given that such fluctuations can be
measured, our Eq. (20) allows one to track the number of
the attached elements at different degrees of stretching
and different temperatures.
On the other hand, in mechanical experiments con-
ducted on single fibers and involving x-ray diffraction
measurements [77], one can, in principle, test the predic-
tion of the model that the configuration with (cid:104)p(cid:105) = 1/2
is the state of maximum disorder. By studying statis-
tics of the observed fluctuations in the steady states, one
can also search for deviations from the static fluctuation-
dissipation relation (5). If found, then they may reveal
the presence of out-of-equilibrium active processes at the
time scales of fast force recovery which would then recon-
cile the mechanical and the biochemical pictures of this
phenomenon [78].
of fluctuations
m. Cooperativity. Statistics
for
groups of myosins has been studied exhausively in
experiments involving active contraction. Considerable
coordination between individual elements was detected,
responsible for synchronized oscillations in close to stall
conditions (our point y = y0) [56, 79–84]. The coopera-
tive behavior was explained by the fact that, due to the
presence of long-range elastic interactions transmited
through compliant backbones,
the mechanical state
of one motor influences the kinetics of other motors
[40, 41]. The implied myosin-myosin coupling was taken
into consideration in models addressing active behavior
of motor groups [85, 86] and emergent phenomena
characterized by large-scale entrainment signatures were
identified [29, 65, 66, 87, 88]. The claims that activity in
such systems is crucial for the emergence of synchronized
oscillations were supported by in vitro assays [65, 66]
showing that the finite size scaling of the fluctuations is
fundamentally differs from the equilibrium one N−1/2.
It is clear that to capture this effect, the original HS
model with rigid backbone and controlled displacement
has to be generalized, but the actual role of activity in
synchronization of cross-bridges is not obvious. Thus,
it has been recently argued [2] that the dominant factor
behind collective behavior is not activity but the long-
range interactions between cross-bridges. The simplest
way to create such "cross-talk" without leaving the HS
framework is to consider the response of a HS system
subjected to a constant force (soft device) rather than
a constant displacement [26].
It has been shown that
in the systems of this type the nonconvexity of the free
energy can resist thermal fluctuations at sufficiently low
temperatures, giving rise to macroscopic cooperativity.
Moreover, in the case of soft and mixed (soft-hard) load-
ings, the pseudocritical point of HS at β = 4 becomes a
real critical point of the Curie type around which fluctu-
ations diverge in the thermodynamic limit and can show
unusual finite-size scaling [2].
V. NONMUSCLE SYSTEMS
The prototypical nature of the HS model makes it rel-
evant outside the skeletal muscle context as well.
In
fact, it can be viewed as a description of a large class
of biological systems involving collectively biased multi-
stable elements and exhibiting, as a result, sigmoidal or
ultrasensitive response at finite temperatures as in our
Fig. 2(a). The HS model describes, perhaps, the most
elementary molecular system capable of transforming in
a Brownian environment a continuous input into a bi-
nary, all-or-none output that is crucial for the fast and
efficient, stroke-type behavior.
We recall that the capacity of multisite systems to flip
in a reversible fashion between several metastable confor-
mations is essential for many processes in cellular phys-
iology, including cell signaling, cell movement, chemo-
taxis, differentiation, and selective expression of genes
[89, 90]. Usually, both the input and the output in such
systems, known as allosteric, are assumed to be of bio-
chemical origin. The HS model, dealing with mechanical
response and relying on mechanical driving, complements
biochemical models and presents a different perspective
on allostery.
n. Hair cell gating. Our first example of hypersen-
sitivity concerns the transduction channels in hair cells
[91]. Each hair cell contains a bundle of N ≈ 50 stere-
ocilia which are mechanically stimulated by the vibra-
tions in the inner ear. The stereocilia possess transduc-
tion channels closed by "gating springs" which can open
(close) in response to a positive (negative) shear strain
X, imposed on the cilia from outside.
The broadly accepted model of this phenomenon [10]
16
views the hair bundle as a set of N bistable springs ar-
ranged in parallel.
It is identical to the HS model if
the folded (unfolded) configurations of cross-bridges are
identified with the closed (opened) states of the channels.
The applied loading, which tilts the potential and biases
in this way the distribution of closed and open configu-
rations, is treated in this model as a hard device of HS.
To stress the equivalence of the results, it is enough
to mention the expression for the total stiffness of a hair
bundle obtained in Ref. [10], which is a direct analog
of our Eq. (15). Moreover, the mechanical experiments,
involving a mechanical solicitation of the hair bundle
through an effectively rigid glass fiber showed that the
stiffness of the hair bundle is negative around the phys-
iological functioning point of the system [11] , which is
fully compatible with the predictions of the HS model.
o. Cell adhesion. A similar analogy can be drawn
between the HS model and the models of collective unzip-
ping for adhesive clusters [9, 92–95]. At the micro-scale
we again deal with N elements representing, for instance,
integrins or cadherins, that are attached in parallel to a
common, relatively rigid pad. The two conformational
states, which can be described by a single spin variable,
are the bound and the unbound configurations.
The binding-unbinding phenomena in a mechanically
biased system of the HS type are usually described by
the Bell model [8], which is a soft device analog of the
HS model with κ0 = ∞. In this model the breaking of
an adhesive bond represents an escape from a metastable
state and the corresponding rates are computed by using
Kramers' theory [95, 96] as in the HS model. In partic-
ular, the rebinding rate is often assumed to be constant
[97, 98], which is also the assumption of HS for the reverse
transition from the post- to the pre-power-stroke state.
More recently, Bell's model was generalized through the
inclusion of ligand tethers, bringing a finite value to κ0
and using the master equation for the probability distri-
bution of attached units [9, 97].
The main difference between the Bell-type models and
the HS model is that the detached state cannot bear
force while the unfolded conformation can. As a result,
while the cooperative folding-unfolding (ferromagnetic)
behavior in the HS model is possible in the soft device
setting [2], similar cooperative binding-unbinding in the
Bell model is impossible because the rebinding of a fully
detached state has zero probablity. To obtain cooperativ-
ity in models of adhesive clusters, one must use a mixed
device, mimicking the elastic backbone and interpolating
between soft and hard driving [9, 26, 95, 99].
p. Synaptic fusion. While muscle tissues maintain
stable architecture over long periods of time, it is feasi-
ble that transitory muscle-type structures can be also as-
sembled to perform particular functions. An interesting
example of such assembly is provided by the SNARE pro-
teins responsible for the fast release of neurotransmitors
from neurons to synaptic clefts. The fusion of synaptic
vesicles with the presynaptic plasma membrane [22, 100]
is achieved by mechanical zipping of the SNARE com-
plexes which can in this way transform from opened to
closed conformation [101].
To complete the analogy, we mention that individual
SNAREs participating in the collective zipping are at-
tached to an elastic membrane that can be mimicked by
an elastic or even rigid backbone [102]. The presence of
a backbone mediating long-range interactions allows the
SNAREs to cooperate in fast and efficient closing of the
gap between the vesicle and the membrane. The analogy
with muscles is corroborated by the fact that synaptic
fusion takes place at the same time scale as the fast force
recovery (1 ms) [103].
q. Macromolecular hairpins. Another class of phe-
nomena that can be rationalized within the HS frame-
work is the ubiquitous flip-flopping of macro-molecular
hairpins subjected to mechanical loading [12–14, 16].
We recall that in a typical experiment of this type, a
folded (zipped) macromolecule is attached through com-
pliant links to micron-sized beads trapped in optical
tweezers. As the distance between the laser beams is
increased, the force applied to the molecule rises up to
a point where the subdomains start to unfold. An indi-
vidual unfolding event may correspond to the collective
rupture of N molecular bonds or an unzipping of a hair-
pin. The corresponding drops in the force accompanied
by an abrupt increase in the total stretch can lead to an
overall negative stiffness response [12, 15, 55].
Realistic examples of unfolding in macromolecules may
involve complex "fracture" avalanches [19] that cannot
be modeled by using the original HS model. However,
the HS theoretical framework is general enough to ac-
commodate hierarchical meta-structures whose stability
can be also biased by mechanical loading. In fact, the
importance of the topology of interconnections among
the bonds and the link between the collective nature of
the unfolding and the dominance of the HS-type parallel
bonding have been long stressed in the studies of protein
folding [104]. The broad applicability of the HS mechan-
ical perspective on collective conformational changes is
also corroborated by the fact that proteins and nucleic
acids exhibit negative stiffness and behave differently in
soft and hard devices [18, 20, 21].
The ensemble dependence in these systems suggests
that additional structural information can be obtained if
the unfolding experiments are performed in the mixed de-
vice setting. The type of loading may be affected through
the variable rigidity of the "handles" [105, 106] or the use
of an appropriate feedback control that can be modeled
in the HS framework by a variable backbone elasticity.
r. Allosteric systems. As we have already men-
tioned, collective conformational changes in distributed
biological systems containing coupled bistable units can
be driven not only mechanically, by applying forces or
displacements, but also biochemically by, say, varying
concentrations or chemical potentials of ligand molecules
in the environment [107]. Such systems can become ul-
trasensitive to external stimulations as a result of the in-
teraction between individual units undergoing conforma-
17
tional transformation which gives rise to the phenomenon
of allostery also known as conformational spread [90,
108]. The switch-like input-output relations are required
in a variety of biological applications because they ensure
both robustness in the presence of external perturbations
and ability to quickly adjust the configuration in response
to selected stimuli [89, 109]. The mastery of control of
biological machinery through mechanically induced con-
formational spread is an important step in designing ef-
ficient biomimetic nanomachines [35, 110–112].
To link this behavior to the HS model, we note that
the amplified dose response, characteristic of allostery, is
analogous to the sigmoidal stress response of the param-
agnetic HS system where an applied displacement plays
the role of the controlled input of a ligand. Usually, in
allosteric protein systems, the ultrasensitive behavior is
achieved as a result of nonlocal interactions favoring all-
or-none types of responses; moreover, the required long-
range coupling is provided by mechanical forces acting
inside membranes and molecular complexes. In the HS
model such coupling is modeled by the parallel arrange-
ment of elements, which preserves the general idea of non-
locality. Despite its simplicity, the appropriately general-
ized HS model [2] captures the main patterns of behavior
exhibited by allosteric systems, including the possibility
of a critical point mentioned in Ref. [107].
VI. CONCLUSIONS
In this paper we presented a perspective on the sem-
inal work of Huxley and Simmons by viewing their re-
sults through the prism of statistical mechanics. This
allowed us to place an emphasis on thermal effects and
equilibrium fluctuations that cannot be ignored in many
biological applications of the HS model.
The chemomechanical approach of HS is based on the
description of a mechanical system with N elements in a
thermal bath in terms of a single deterministic reaction
equation. Instead, our analysis starts with the analogy
between the HS model in a hard device and the param-
agnetic Ising model where the average conformation of
a cross-bridge viewed as the counterpart of magnetiza-
tion. In view of this analogy, the HS model describes a
size indifferent mean-field statistical mechanical system
which explains why the many-body stochastic dynamics
can be modeled by a single chemical reaction. The anal-
ogy with paramagnetism is, however, not complete, as is
revealed by the phenomena of negative susceptibility and
pseudo-criticality that we identify with the HS model. In
particular, we show that while genuine criticality requires
cooperativity, in the HS model the collective response is
imposed through the rigid backbone rather than being
an emergent property.
Some of the most interesting findings of this paper con-
cern the thermal properties of the HS system. Thus,
our analysis highlights the previously unnoticed temper-
ature robustness of the pseudocritical state where specific
heat vanishes and fluctuations become temperature inde-
pendent. We also quantified the temperature variations
during fast unfolding and demonstrated that isothermal
and adiabatic responses may differ. These observations
point towards the importance of the nonorthodox experi-
mental protocols combining mechanical and calorimetric
measurements. In particular, the revealed fluctuation de-
pendence of equilibrium susceptibilities suggests a non-
crystallographic way for the evaluation of the number of
folding elements.
To account for fluctuations in kinetics we, following
Refs. [29, 30], went beyond the reaction-based modeling
of the averages and studied the time dependence of the
probability distributions for various parameters during
mechanical transients. The results of the deterministic
model of HS are, of course, recovered from the analysis of
the evolution of the first moments of these distributions.
While being the simplest mean-field description of
a broad class of biological phenomena, the HS model
clearly misrepresents the important elastic coupling be-
tween the folding elements, as has been long realized in
muscle physiology [40, 41]. This drawback can be reme-
died by taking into account the mechanical feedback in-
duced by the backbone elasticity [2] resulting in various
18
degrees of synchronization already at zero temperature
[26]. A more important limitation of the HS model is the
neglect of the ATP-fueled activity which can crucially in-
terfere with passive folding [29, 65, 66, 85–88, 113]. The
account of nonthermal driving in the HS setting produces
new qualitative effects [78, 114] which opens the possibil-
ity to build a fully mechanistic analog of the enzimatic
cycle originating in the work of Lymn and Taylor. Quan-
titative applications of the HS model in and outside the
muscle context also call upon the account of complex
geometry, hierarchical architecture, soft-spin-type multi-
stability, and short-range interactions. All these poten-
tial augmentations, however, will not diminish the role of
the original HS model as a source of fundamental physical
intuition about the behavior of a wide class of biological
systems.
VII. ACKNOWLEDGMENTS
The authors thank R. Sheshka, P. Recho, T. Leli`evre
and J.-M. Allain and the anonymous reviewers for helpful
comments.
[1] A. F. Huxley and R. M. Simmons, Nature 233, 533
[18] T. Bornschlogl and M. Rief, Phys. Rev. Lett. 96, 118102
(1971).
(2006).
[2] M. Caruel, J. M. Allain, and L. Truskinovsky, Phys.
[19] A. Srivastava and R. Granek, Phys. Rev. Lett. 110,
Rev. Lett. 110, 248103 (2013).
138101 (2013).
[3] T. L. Hill, Prog. Biophys. Molec. Biol. 29, 105 (1976).
[4] A. F. Huxley and S. Tideswell, J. Muscle Re. Cell M.
[20] U. Gerland, R. Bundschuh, and T. Hwa, Biophysj 84,
2831 (2003).
17, 507 (1996).
[21] N. Thomas and Y. Imafuku, J Theor Biol 312, 96
[5] G. Piazzesi and V. Lombardi, Biophys. J. 68, 1966
(2012).
(1995).
[6] D. Smith, M. Geeves, J. Sleep, and S. Mijailovich, Ann.
[22] T. C. Sudhof, Neuron 80, 675 (2013).
[23] L. L. Bonilla, A. Carpio, A. Prados, and R. R. Rosales,
Biomed. Eng. 36, 1624 (2008).
Phys. Rev. E 85, 031125 (2012).
[7] M. Linari, M. Caremani, and V. Lombardi, P Roy. Soc.
Lond. B Bio. 277, 19 (2010).
[8] G. Bell, Science 200 (1978).
[9] T. Erdmann and U. S. Schwarz, Eur. Phys. J. E 22, 123
(2007).
[10] J. Howard and A. J. Hudspeth, Neuron 1, 189 (1988).
[11] P. Martin, A. Mehta, and A. Hudspeth, Proc. Natl.
Acad. Sci. U.S.A. 97, 12026 (2000).
[12] J. Liphardt, B. Onoa, S. B. Smith, I. Tinoco,
and
C. Bustamante, Science 292, 733 (2001).
[13] N. Bosaeus, A. H. El-Sagheer, T. Brown, S. B. Smith,
and B. Norden, Proc.
B. Akerman, C. Bustamante,
Natl. Acad. Sci. U.S.A. 109, 15179 (2012).
[14] M. T. Woodside, C. Garcia-Garcia, and S. M. Block,
Curr. Opin. Chem. Biol. 12, 640 (2008).
[15] A. N. Gupta, A. Vincent, K. Neupane, H. Yu, F. Wang,
[24] T. L. Hill, Prog. Biophys. Molec. Biol. 28, 267 (1974).
[25] L. Marcucci and L. Truskinovsky, Phys. Rev. E 81
(2010).
[26] M. Caruel, J.-M. Allain, and L. Truskinovsky, J. Mech.
Phys. Solids 76, 237 (2015).
[27] S. Pradhan, A. Hansen, and B. K. Chakrabarti, Rev.
Mod. Phys. 82, 499 (2010).
[28] R. Balian, From Microphysics to Macrophysics, Meth-
ods and Applications of Statistical Physics (Springer
Science & Business Media, 2006).
[29] A. Vilfan and T. Duke, Biophys. J. 85, 818 (2003).
[30] T. Erdmann, P. J. Albert, and U. S. Schwarz, J. Chem.
Phys. 139, 175104 (2013).
[31] J. S. Rowlinson, Liquids and liquid mixtures, Modern as-
pects series of chemistry (Butterworths, London, 1959).
[32] D. R. Squire, A. C. Holt, and W. G. Hoover, Physica
and M. T. Woodside, Nat Phys 7, 631 (2011).
42, 388 (1969).
[16] A. Prados, A. Carpio, and L. L. Bonilla, Phys. Rev. E
86, 021919 (2012).
[33] J. F. Lutsko, J. Appl. Physiol. 65, 2991 (1989).
[34] J. P. Wittmer, H. Xu, P. Poli´nska, F. Weysser, and
[17] V. Munoz, E. R. Henry, J. Hofrichter, and W. A. Eaton,
J. Baschnagel, J. Chem. Phys. 138, 191101 (2013).
Proc. Natl. Acad. Sci. U.S.A. 95, 5872 (1998).
[35] Z. Wu, R. Harne, and K. Wang, J. Intell. Mat. Syst.
Struct. 27, 1189 (2016).
19
[36] E. Eisenberg and T. Hill, Prog. Biophys. Molec. Biol.
33, 55 (1978).
[37] L. E. Ford, A. F. Huxley, and R. M. Simmons, J. Phys-
iol. 269, 441 (1977).
[65] L. Hilbert, S. Cumarasamy, N. B. Zitouni, M. C.
Mackey, and A.-M. Lauzon, Biophysj 105, 1466 (2013).
[66] L. Hilbert, Z. Balassy, N. B. Zitouni, M. C. Mackey,
and A.-M. Lauzon, Biophys. J. 108, 622 (2015).
[38] L. E. Ford, A. F. Huxley, and R. M. Simmons, The
[67] J. E. Baker, C. Brosseau, P. B. Joel, and D. M. War-
Journal of physiology 311, 219 (1981).
shaw, Biophysj 82, 2134 (2002).
[39] H. Kojima, A. Ishuima, and T. Yanagida, Proc. Natl.
[68] J. E. Baker, C. Brosseau, P. Fagnant, and D. M. War-
Acad. Sci. U.S.A. 91, 12962 (1994).
shaw, J. Biol. Chem. 278, 28533 (2003).
[40] K. Wakabayashi, Y. Sugimoto, H. Tanake, Y. Ueno,
Y. Takezawa, and Y. Amemiya, Biophys. J. 67, 2422
(1994).
[69] B. D. Haldeman, R. K. Brizendine, K. C. Facemyer,
J. E. Baker, and C. R. Cremo, J. Biol. Chem. 289,
21055 (2014).
[41] H. Huxley, A. Stewart, H. Sosa, and T. Irving, Biophys.
[70] Anneka M Hooft, Erik J Maki, K. K. Cox, and J. E.
J. 67, 2411 (1994).
[42] G. Piazzesi, M. Reconditi, M. Linari, L. Lucii,
P. Bianco, E. Brunello, V. Decostre, A. Stewart, D. B.
Gore, T. C. Irving, M. Irving, and V. Lombardi, Cell
131, 784 (2007).
[43] M. Irving, G. Piazzesi, L. Lucii, Y. Sun, J. Harford,
I. Dobbie, M. Ferenczi, M. Reconditi, and V. Lombardi,
Nat. Struct. Biol. 7, 482 (2000).
[44] A. V. Hill, P Roy. Soc. Lond. B Bio. 126, 136 (1938).
[45] G. Piazzesi, A. Reconditi, N. Koubassova, V. Decostre,
M. Linari, L. Lucii, and V. Lombardi, J. Physiol. 549,
93 (2003).
[46] V. Decostre, P. Bianco, V. Lombardi, and G. Piazzesi,
Proc. Natl. Acad. Sci. USA 102, 13927 (2005).
[47] K. W. Ranatunga, M. E. Coupland, G. J. Pinniger,
H. Roots, and G. W. Offer, J. Physiol. 585, 263 (2007).
[48] M. Coupland and K. W. Ranatunga, J. Physiol. 548,
439 (2003).
Baker, Biochemistry 46, 3513 (2007).
[71] J. Del R Jackson and J. E. Baker, Physical Chemistry
Chemical Physics 11, 4808 (2009).
[72] M. Caremani, L. Melli, M. Dolfi, V. Lombardi, and
M. Linari, J. Physiol. 593, 3313 (2015).
[73] S. H. Gilbert and L. E. Ford, Biophysj 54, 611 (1988).
and R. Cooke,
[74] E. Pate, G. J. Wilson, M. Bhimani,
Biophysj 66, 1554 (1994).
[75] M. Linari, R. Bottinelli, M. Pellegrino, M. Reconditi,
C. Reggiani, and V. Lombardi, J. Physiol. - London
555, 851 (2004).
[76] K. W. Ranatunga, The Journal of physiology 588, 3657
(2010).
[77] M. Reconditi, Rep. Prog. Phys. 69, 2709 (2006).
[78] R. Sheshka and L. Truskinovsky, Phys. Rev. E 89,
012708 (2014).
[79] A. F. Huxley, J. Physiol. 243, 1 (1974).
[80] C. F. Armstrong, A. F. Huxley, and F. J. Julian, J.
[49] R. C. Woledge, C. J. Barclay, and N. A. Curtin, P Roy.
Physiol. 186, 22P (1966).
Soc. Lond. B Bio. 276, 2685 (2009).
[50] M. Linari and R. C. Woledge, J. Physiol. 487, 699
[81] K. A. P. Edman, J. Physiol. - London 404, 301 (1988).
[82] K. Yasuda, Y. Shindo, and S. Ishiwata, Biophys. J. 70,
(1995).
1823 (1996).
[51] C. Veigel, M. Bartoo, D. White, J. Sparrow, and J. Mol-
[83] K. A. P. Edman and N. Curtin, J. Physiol. - London
loy, Biophys. J. 75, 1424 (1998).
534, 553 (2001).
[52] M. Tyska and D. Warshaw, Cell Motil. Cytoskel. 51, 1
[84] P. Y. Pla¸cais, M. Balland, T. Guerin, J. F. Joanny, and
(2002).
[53] E. Brunello, M. Caremani, L. Melli, M. Linari,
M. Fernandez-Martinez, T. Narayanan, M.
Irving,
G. Piazzesi, V. Lombardi, and M. Reconditi, The Jour-
nal of physiology 592, 3881 (2014).
[54] N. van Kampen, Stochastic Processes in Physics and
P. Martin, Phys. Rev. Lett. 103 (2009).
[85] T. A. J. Duke, Proc. Natl. Acad. Sci. USA 96, 2770
(1999).
[86] T. Duke, Philos. T. Roy. Soc. B 355, 529 (2000).
[87] M. Badoual, F. Julicher, and J. Prost, Proc. Natl. Acad.
Sci. U.S.A. 99, 6696 (2002).
Chemistry (North-Holland, Amsterdam, 2001).
[88] S. Walcott and S. X. Sun, Physical Chemistry Chemical
[55] J.-D. Wen, M. Manosas, P. T. X. Li, S. B. Smith,
C. Bustamante, F. Ritort, and I. Tinoco, Biophys. J.
92, 2996 (2007).
[56] R. Podolsky, Nature 188, 666 (1960).
[57] M. Civan and R. Podolsky, J. Physiol. 184, 511 (1966).
[58] R. Podolsky, A. C. Nolan, and S. A. Zaveler, Proc.
Natl. Acad. Sci. U.S.A. 64, 504 (1969).
[59] A. F. Huxley and R. M. Simmons, J. Physiol. 208, 52P
(1970).
Physics 11, 4871 (2009).
[89] T. A. Duke, N. Le Novere, and D. Bray, J. Mol. Biol.
308, 541 (2001).
[90] D. Bray, Journal of Molecular Biology 425, 1410 (2013).
[91] V. Bormuth, J. Barral, J. F. Joanny, F. Julicher, and
P. Martin, Proc. Natl. Acad. Sci. U.S.A. 111, 7185
(2014).
[92] B. Chen and H. Gao, Biophys. J. 101, 396 (2011).
[93] H. Yao and H. Gao, J. Mech. Phys. Solids 54, 1120
[60] K. Holmes and M. Geeves, Philos. T. Roy. Soc. B 355,
(2006).
419 (2000).
[94] H. Gao, J. Qian, and B. Chen, J. R. Soc. Interface 8,
[61] I. Rayment, H. Holden, M. Whittaker, C. Yohn,
M. Lorenz, K. Holmes, and R. Milligan, Science 261,
58 (1993).
1217 (2011).
[95] U. S. Schwarz and S. A. Safran, Reviews of Modern
Physics 85, 1327 (2013).
[62] R. Dominguez, Y. Freyzon, K. M. Trybus, and C. Co-
[96] P. Hanggi, P. Talkner, and M. Borkovec, Reviews of
hen, Cell 94, 559 (1998).
[63] R. Lymn and E. Taylor, Biochemistry 10, 4617 (1971).
[64] B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts,
and P. Walter, Molecular Biology of the Cell, 5th ed.
(Garland Science, New York, 2007).
Modern Physics 62, 251 (1990).
[97] T. Erdmann and U. S. Schwarz, Phys. Rev. Lett. 92,
108102 (2004).
[98] T. Erdmann and U. S. Schwarz, J. Chem. Phys. 121,
8997 (2004).
20
[99] U. Seifert, Phys. Rev. Lett. 84, 2750 (2000).
[107] J. P. Changeux, J. Thi´ery, Y. Tung, and C. Kittel,
[100] T. C. Sudhof and J. E. Rothman, Science 323, 474
Proc. Natl. Acad. Sci. USA 57, 335 (1967).
(2009).
[108] D. Bray and T. Duke, Annu Rev Biophys Biomol Struct
[101] D. Min, K. Kim, C. Hyeon, Y. H. Cho, Y.-K. Shin, and
33, 53 (2004).
T.-Y. Yoon, Nat. Commun. 4, 1705 (2013).
[109] B. M. Slepchenko and M. Terasaki, Current Opinion in
[102] F. Li, F. Pincet, E. Perez, W. S. Eng, T. J. Melia, J. E.
Rothman, and D. Tareste, Nature Structural & Molec-
ular Biology 14, 890 (2007).
[103] S. Zorman, A. A. Rebane, L. Ma, G. Yang, M. A.
Molski, J. Coleman, F. Pincet, J. E. Rothman, and
Y. Zhang, Elife 3, e03348 (2014).
[104] H. Dietz and M. Rief, Phys. Rev. Lett. 100, 098101
(2008).
Genetics & Development 14, 428 (2004).
[110] O. Miyashita, J. N. Onuchic, and P. G. Wolynes, PNAS
100, 12570 (2011).
[111] B. Yurke, A. J. Turberfield, A. P. Mills, F. C. Simmel,
and J. L. Neumann, Nature 406, 605 (2000).
[112] R. L. Harne, Z. Wu, and K. W. Wang, J. Mech. Des
138, 021402 (2016).
[113] T. Guerin, J. Prost, and J. F. Joanny, Eur. Phys. J.
[105] R. Cerf, Proc. Natl. Acad. Sci. U.S.A. 75, 2755 (1978).
[106] E. Pfitzner, C. Wachauf, F. Kilchherr, and B. Pelz,
34, 60 (2011).
[114] Sheshka, R, Recho, P, and Truskinovsky, Lev, Phys.
Angew. Chem. Int. Ed. 52, 335 (2013).
Rev. E 93, 052604 (2016).
|
1606.06594 | 1 | 1606 | 2016-06-20T00:34:03 | Comment on Mente, O'Donnell, Rangarajan, et al. "Associations of urinary sodium excretion with cardiovascular events in individuals with an without hypertension: a pooled analysis of data from four studies" | [
"physics.bio-ph",
"q-bio.QM"
] | Mente, O'Donnell, Rangarajan, et al. ignore a possible source of bias that may invalidate their finding of an anticorrelation between sodium intake and cardiovascular events | physics.bio-ph | physics |
Comment on Mente, O'Donnell, Rangarajan,
et al. "Associations of urinary sodium
excretion with cardiovascular events in
individuals with and without hypertension: a
pooled analysis of data from four studies"
Dept. Physics, Washington University St. Louis, Mo. 63130
J. I. Katz
September 3, 2018
Abstract
Mente, O'Donnell, Rangarajan, et al. ignore a possible source of
bias that may invalidate their finding of an anticorrelation between
sodium intake and cardiovascular events.
Mente, O'Donnell, Rangarajan, et al. [1] ignore a possible source of bias
in their analysis of association of sodium excretion with cardiovascular events.
The population at any level of blood pressure will comprise individuals with
varying levels of sodium sensitivity, and those with higher sensitivity will
achieve that blood pressure if they ingest less sodium. As a result, within
any blood pressure level, for example, normotension, there will be an anti-
correlation between sodium intake and sodium sensitivity.
If sodium sensitivity is causal of cardiovascular events or death this could
explain the observational anticorrelation between sodium intake and these
events. Such a causal relation between sodium sensitivity and adverse out-
comes is plausible: sensitive individuals are likely to have been hypertensive
in the past and to be so in the future. As a result, the observational associ-
ation of low sodium excretion with cardiovascular events may not be causal,
but an artefact of the anticorrelation between sodium intake and sensitivity
within the normotensive group.
1
References
[1] Mente A, O'Donnell M, Rangarajan S, et al. Associations of urinary
sodium excretion with cardiovascular events in individuals with and with-
out hypertension: a pooled analysis of data from four studies Lancet on
line 20 May 2016
2
|
1808.04499 | 1 | 1808 | 2018-08-14T01:00:30 | Dynamical and Coupling Structure of Pulse-Coupled Networks in Maximum Entropy Analysis | [
"physics.bio-ph",
"q-bio.NC"
] | Maximum entropy principle (MEP) analysis with few non-zero effective interactions successfully characterizes the distribution of dynamical states of pulse-coupled networks in many experiments, e.g., in neuroscience. To better understand the underlying mechanism, we found a relation between the dynamical structure, i.e., effective interactions in MEP analysis, and the coupling structure of pulse-coupled network to understand how a sparse coupling structure could lead to a sparse coding by effective interactions. This relation quantitatively displays how the dynamical structure is closely related to the coupling structure. | physics.bio-ph | physics | Dynamical and Coupling Structure of Pulse-Coupled Networks in Maximum Entropy Analysis
Zhi-Qin John Xu1,∗ Douglas Zhou2,† and David Cai1,2,3
1NYUAD Institute, New York University Abu Dhabi,
Abu Dhabi, United Arab Emirates,
2School of Mathematical Sciences,
MOE-LSC and Institute of Natural Sciences,
Shanghai Jiao Tong University, Shanghai, P.R. China,
3Courant Institute of Mathematical Sciences and Center for Neural Science,
New York University, New York, New York, USA.
8
1
0
2
g
u
A
4
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
9
9
4
4
0
.
8
0
8
1
:
v
i
X
r
a
(Dated: August 15, 2018)
Maximum entropy principle (MEP) analysis with few non-zero effective interactions successfully character-
izes the distribution of dynamical states of pulse-coupled networks in many experiments, e.g., in neuroscience.
To better understand the underlying mechanism, we found a relation between the dynamical structure, i.e., ef-
fective interactions in MEP analysis, and the coupling structure of pulse-coupled network to understand how
a sparse coupling structure could lead to a sparse coding by effective interactions. This relation quantitatively
displays how the dynamical structure is closely related to the coupling structure.
PACS numbers: 89.70.Cf, 87.19.lo, 87.19.ls, 87.19.ll
Binary-state networks -- each node in one sampling time bin
is binary-state -- arise from many research fields, e.g., gene
regulatory modeling and neural dynamics [15, 23, 26]. Sta-
tistical distributions of network states are essential to encode
information [6, 11, 18, 20, 25]. For example, with statisti-
cal distributions of network states, experimental studies show
that rats can perform awake replays of remote experiences
in hippocampus [10]. Many works effectively characterize
the distribution of 2n network states for n binary-state nodes
in various systems, e.g., a network of ∼ 100 neurons [8],
with a low-order maximum entropy principle (MEP) analy-
sis [2, 4, 13, 14, 19, 22, 24, 27] -- a method with few (far less
than 2n) non-zero effective interactions (see a precise defini-
tion in Eq. (1)) constrained by low-order statistics. We can
then regard those effective interactions as a sparse coding of
the information that encoded in the state distribution. To un-
derstand coding schemes of network systems, it is important,
however, yet to understand what leads to the sparsity of effec-
tive interactions. In this work, we would mainly use neural
networks as examples for illustration, while our results apply
to general binary-state networks.
Estimated by dynamical data of a network system, effec-
tive interactions reflect a dynamical structure of the network.
This dynamical structure has been used to study the functional
connectivity of networks [7, 27]. For example, experimental
studies show that the second-order effective interaction map
of the retina is sparse and dominated by local overlapping ef-
fective interaction modules [7]. Network dynamical structure
often closely relates to the underlying coupling structure [29].
For example, when the input of each node is independent to
others, i) high-order (≥ 2) effective interactions are zero in
a network of no connections, ii) high-order effective interac-
tions are large in a dense and strong connected excitatory net-
work. To efficiently encode information, a realistic system
∗ [email protected]
† [email protected]
often incorporates a coupling structure with certain features
[3, 17], e.g., sparsity, small-world, or scale-free. However, it
is still unclear how the coupling structure affects the dynami-
cal structure of effective interactions.
In this letter, we consider a general class of pulse-coupled
networks. The state of each node is binary-state, i.e., ac-
tive when the node sends pulses to its child nodes, otherwise,
silent. We observed a Fact that leads to an explicit relation --
which is independent of node dynamics -- between the cou-
pling structure and the number of non-zero effective interac-
tions in the full-order MEP analysis (constrained by all mo-
ments). We examine our observed Fact by numerical simula-
tions. Through our analysis, we can estimate an upper bound
of the number of non-zero effective interactions for a given
coupling structure when the external input of each node is in-
dependent with each other. Our results show that a sparse
network could lead to a lot of vanishing high-order effective
interactions. For illustration, we estimate the number of non-
zero effective interactions for each order in a network with
Erdos-Renyi connection structure, in which our estimation is
much smaller than Ck
n, the number of all possible kth-order
effective interactions. Our results establish a connection be-
tween the dynamical structure and the network coupling struc-
ture. This connection provides an insight into how a sparse
coupling structure can lead to a sparse coding scheme.
In the following analysis, we use binary vector V (l) =
(σ1,··· ,σn) ∈ {0,1}n to represent the state of n nodes within
the sampling time bin labeled by l. To obtain correlations up
to the mth-order requires to evaluate all (cid:104)σi1 ···σiM(cid:105)E, where
1 ≤ i1 < i2 < ··· < iM ≤ n, 1 ≤ M ≤ m, and (cid:104)·(cid:105)E is de-
fined by(cid:104)g(l)(cid:105)E = ∑NT
l=1 g(l)/NT for any function g(l) and NT
is the total number of sampling time bins in the recording.
The mth-order MEP analysis is to find the desired probabil-
ity distribution P(V ) for n nodes by maximizing the entropy
S ≡ −∑V P(V )logP(V ) subject to correlations up to the mth-
order (m ≤ n). Then, the unique distribution can be solved
as
Pm(V ) =
1
Z
exp(
m
∑
k=1
n
∑
i1<···<ik
Ji1···ikσi1 ···σik ),
(1)
where, following the terminology of statistical physics, we
call Ji1···ik a kth-order effective interaction (1 ≤ k ≤ m), the
partition function Z is the normalization factor. Eq.
(1) is
referred to as the mth-order MEP distribution.
First, we discuss the relationship between effective inter-
actions and the statistical distribution of network states. By
(1) for Pn(V ), we
taking logarithm of both sides of Eq.
can get a set linear equations of all-order effective interac-
tions for all states V . Since Pn is the same as the exper-
imental observed distribution [1], we can obtain the effec-
tive interactions in Pn in terms of the experimental observed
distribution [28]. For example, n = 3, we can obtain J1 =
log(P100/P000) and J12 = log(P110/P010)− J1, where Pσ1σ2σ3
represents the probability of the network state (σ1,σ2,σ3).
By applying P(σ1,σ2,σ3) = P(σ1σ2,σ3)P(σ2,σ3), we have
P(σ1=0σ2=0,σ3=0) and J12 = log P(σ1=1σ2=1,σ3=0)
J1 = log P(σ1=1σ2=0,σ3=0)
P(σ1=0σ2=1,σ3=0) −
1 −J1. Our earlier study has shown a recursive structure
J1 (cid:44) J1
among effective interactions, that is, the (k + 1)st-order ef-
fective interaction J123...(k+1) can be obtained as follows [28]:
First, we switch the state of the (k + 1)st node in J123...k from
silent to active to obtain a new term J1
123...k, e.g., from J1 to
J1
1 ; Then, we subtract J123...k from the new term to obtain
J123...(k+1), i.e.,
J123...(k+1) = J1
123...k − J123...k.
(2)
Without lost of generality, we randomly select two nodes la-
beled by 1 and 2. By the recursive relation, any kth-order ef-
fective interaction that includes node 1 and 2 can be expressed
as the summation of terms with the following basic form
.
12(σ3,··· ,σn) = log
Jb
− log
P(σ1 = 1σ2 = 1,σ3,··· ,σn)
P(σ1 = 0σ2 = 1,σ3,··· ,σn)
P(σ1 = 1σ2 = 0,σ3,··· ,σn)
P(σ1 = 0σ2 = 0,σ3,··· ,σn)
12(1,0,··· ,0) − Jb
(3)
12(0,0,··· ,0) and
For example, J123 = Jb
12(0,1,0,··· ,0)]− J123. We can
J1234 = [Jb
observe that if nodes 1 and 2 are independent conditioned on
all other nodes, i.e., P(σ1σ2 = 1,σ3,··· ,σn) = P(σ1σ2 =
0,σ3,··· ,σn), any effective interaction containing these two
nodes is zero.
12(1,1,0,··· ,0)− Jb
Next, we would show what kind of coupling structure could
entail the conditional independence of two nodes. Here, we
define some notations. In any sampling time bin [0,∆) with
state V = (σ1,··· ,σn), ∀t ∈ [0,∆), we denote Ii,t as node i's
input from the outside of the network, denote wi j(t) as the
input from the node i to node j, denote C(i) as the set of all
child notes of node i, denote Ui = C(i)∪{i}, denote P(e) as
the probability of event e, denote U0 = {1,2,··· ,n}.
Fact. For n pulse-coupled nodes with binary-state dynam-
ics on a network with a coupling structure G0, in any sam-
pling time bin [0,∆), ∀t ∈ [0,∆), ∀i1, j1 ∈ U0, we assume that:
2
(a) the external inputs of each node are independent to oth-
ers, i.e., P(Ii1,t ,Ij1,t ) = P(Ii1,t )P(Ij1,t ); (b) whether a parent
node sends spikes to its child nodes only depends on its state,
i.e., P(wi1 j1(t),V ) = W (σi1,i1, j1,t), where W (·,·,·,·) is a real
function. ∀i, j ∈ U0, if they neither are connected nor share
any common child node, i.e., Ui ∩Uj = φ, then, node i and j
are independent conditioned on the state of all other nodes,
i.e.,
P(σi,σ jH) = P(σiH)P(σ jH),
where H is a possible state of nodes in U0\{i, j}.
(4)
We justify our two assumptions as follows. To avoid the in-
fluence of correlation in external inputs when we are studying
the relation between the dynamical structure and the coupling
structure, we assume that the external input of each node is in-
dependent to others, i.e., assumption (a). The second assump-
tion implicates a Markov-like property; that is, for a connected
pair of pulse-coupled nodes in an equilibrium state, the pulse
from the parent node to the child node only depends on the
state of the parent node but is independent of inputs imposed
on the parent node. For example, in neural networks, a neuron
sends out spikes only when this neuron is active, regardless of
what inputs are imposed on the neuron.
The argument for the conclusion in Eq. (4) is as follows.
By assumption (a), node i and node j can be dependent only
through the coupling structure G0. When we are considering
how node i and node j affect each other by changing their
states through the coupling structure G0, we can consider a
simplified coupling structure, G1, which ignores those con-
nections that are independent of states of node i and node j,
i.e., σi and σ j. ∀k ∈ Uo\{i, j}, i.e., any other node k, its state
σk is fixed when we are considering the conditional probabil-
ity in Eq. (4). By assumption (b), for node k's any child node
l, the input from node k to node l is independent of σi and σ j.
Thus, the connections started from those nodes in Uo\{i, j}
are fixed for different states of σi and σ j. Therefore, G1 is
a simplified coupling structure that only keeps those connec-
tions originated from node i and node j in G0. In G1, any con-
nection only exists in either sub-network Ui or sub-network
Uj. Under the condition Ui ∩ Uj = φ, i.e., they neither are
connected nor share any common child node, sub-network Ui
and sub-network Uj are two isolated sub-networks. σi and σ j
cannot affect each other by changing their states through the
coupling structure G1, that is, node i and j are independent
conditioned on the states of all other nodes.
Fig.1 displays an example to illustrate our observed Fact.
The coupling structure G0 is shown in Fig.1a. We focus on
node 1 and node 2, where they neither are connected nor share
any child node. When the state of other nodes (black) are
fixed, all outputs from black nodes can be ignored in the sim-
plified coupling structure G1, as shown in Fig.1b. Node 1
and node 2 respectively belong to two separate sub-networks.
Therefore, nodes 1 and node 2 are independent conditioned
on the state of all other nodes.
Based on the recursive structure of effective interactions
and the observed Fact, we reach the following conclusion:
with the two assumptions in the observed Fact, for a group of
3
analysis Pn for this ring network. As shown in Fig.2c, the ef-
fective interaction strengths of independent pairs (J24 and J13)
are within the statistical error of shuffled results (red). Since
every high-order (≥ 3) effective interaction includes at least
one independent pair of nodes, as predicted, the strengths of
all high-order effective interactions are within the statistical
error of shuffled results as shown in Fig.2d.
The second example in the second row in Fig.2, results
are similar that dependent pairs and independent pairs can
be identified through our observed Fact, and the strength of
any effective interaction that includes the independent pair of
nodes (node 1 and node 3) is within the statistical error of
shuffled data. In this example, J124 is very small, i.e., within
the statistical error of shuffled results. However, in our estima-
tion by our conclusion, we do not categorized J124 to the class
of zero-strength effective interactions. This example indicates
that we estimate an upper bound of the number of non-zero ef-
fective interactions. For a network of all excitatory nodes with
the same coupling structure as the one in Fig.1e, J124 is sig-
nificantly larger than zero (not shown). Since the strength of
high-order effective interactions is small, a very long record-
ing constraints us from examining ∆i j(H) for a large network.
Base on the relation between the coupling structure and ef-
fective interactions, the number of non-zero high-order effec-
tive interactions can be small in a sparse connected network
compared with Ck
n, which is the number of all possible kth-
order interactions. For example, we estimate the number of
each-order non-zero effective interactions in a network with
an Erdos-Renyi connection structure. We randomly generate
1000 networks of 100 nodes with an Erdos-Renyi connection.
The connection probability between two nodes is 0.05. As
shown in Fig.3, the number of non-zero kth-order (k > 1) ef-
fective interactions is much smaller than Ck
100 (too large to be
shown). The number of high-order effective interactions (or-
der higher than 11th) almost vanishes (order higher than 20th
not shown).
In summary, we have established a relation between effec-
tive interactions in MEP analysis and the coupling structure of
pulse-coupled networks to understand how a sparse coupling
structure could lead to a sparse coding by effective interac-
tions. This relation quantitatively displays how the dynamical
structure closely relates to the coupling structure.
Even though high-order effective interactions are often
much smaller compared with low-order ones [28], it is still
unclear why small high-order effective interactions do not ac-
cumulate to have a significant effect in a large network [8, 21].
For example, MEP distribution with a sparse low-order effec-
tive interactions -- non-zero effective interactions are sparse
and vanish when the order is high than the eighth-order -- can
well capture the state distribution of 99 ganglion cells in the
salamander retina responding to a natural movie clip or natural
pixel [8]. In this study, we show that a large amount of effec-
tive interactions vanish in a sparse coupling structure; thus,
rationalizing the absence of the accumulation of high-order
interactions for a large network.
Finally, we point out that some important issues remain to
be elucidated in the future. First, we have ignored correlations
in external inputs when estimating the number of non-zero
(a)
(b)
FIG. 1: Structure v.s. Simplified Structure.
nodes {i1,i2,··· ,ik}, if there exists at least one pair of nodes
that neither are connected nor share any child node, effective
interaction Ji1,i2,··· ,ik is zero.
In the system we would use to examine our conclusion is
an integrate-and-fire (I&F) network, a general pulse-coupled
network, with both excitatory and inhibitory nodes [29]. For
the ith node, the dynamics of its state variable xi with time
scales τ is governed by
− (gbg
i + gex)(xi − xex)− gin
i (xi − xin),
(5)
xi = −xi
τ
j,k)/σ in]
j,k (T in
i j) in gex
i
i = ∑ j ∑k Sex
jth excitatory nodes, and gin
where xex and xin are the reversal values of excitation
i = f ∑k H(t −
gbg
(ex) and inhibition (in), respectively.
i,k)exp[−(t − T F
T F
i,k)/σ ex] is the background input with mag-
nitude f and time scale σ ex, T F
i,k is a Poisson process with
rate µ, H(·) is the Heaviside function, gex
i j H(t −
j,k)exp[−(t − T ex
j,k)/σ ex] is the excitatory pulse effective
T ex
interaction from other
i =
j,k)exp[−(t − T in
i jH(t − T in
∑ j ∑k Sin
is the inhibitory
pulse effective interaction from other jth inhibitory nodes.
The jth excitatory (inhibitory) node x j evolves continuously
(5) until it reaches a firing threshold xth.
according to Eq.
That moment in time is referred to as a firing event (say, the
kth spike) and denoted by T ex
j,k). Then, x j is reset to the
reset value xr (xin < xr < xth < xex) and held xr for an absolute
refractory period of τref. Each spike emerging from the jth
excitatory (inhibitory) node causes an instantaneous increase
Sex
i j (Sin
i j are the excitatory and
inhibitory coupling strengths, respectively. The model (5) de-
scribes a general class of physical networks [5, 9, 15, 26, 29].
The first example, two excitatory and two inhibitory I&F
nodes form a ring coupling structure (Fig.2a). For any pair of
nodes, say, node i and j, we compute ∆i j(H) = P(σi = 1σ j =
1,H) − P(σi = 1σ j = 0,H), where H is one state of other
two nodes. By our observed Fact, the conditional indepen-
dent pairs are (neuron 1,neuron 3) and (neuron 2,neuron 4),
and other pairs are categorized as dependent pairs. In Fig.2b,
the strengths of ∆i j(H) of independent pairs (green) are almost
two orders of magnitude smaller than those of dependent pairs
(red). We then shuffle spike trains of each node. We similarly
compute ∆i j(H) for 10 different shuffled data. Blue dots and
cyan dots in Fig.2b are results of all shuffled data of depen-
dent pairs and independent pairs, respectively. The strength of
∆i j(H) of independent pairs (green) -- computed from the ob-
served data -- are within the statistical error of shuffled data.
We then solve effective interactions in the full-order MEP
i ), where Sex
i j and Sin
(gin
4
(a)
(e)
(b)
(c)
(d)
(f)
(g)
(h)
FIG. 2: Structure v.s. effective interactions of I&F networks . Each row shows a numerical case. In the first column, black
arrows and red arrows represents excitatory and inhibitory connections, respectively. In the second column, red and green dots
are the strengths of ∆i j(H) of dependent and independent pairs, respectively. Blue dots and cyan dots are the strengths of
∆i j(H) of dependent and independent pairs from ten shuffled spike trains, respectively. The third and fourth columns display
absolute effective interaction strengths (blue bars). The corresponding node indexes for each effective interaction are shown in
the abscissa. The mean and standard deviation of absolute strengths of each effective interaction of ten shuffled spike trains are
also displayed by garnet bars. The simulation time for each network is 1.2× 108 ms. The time bin size for analysis is 10ms
[22, 24]. Independent Poisson inputs for each network are µ = 0.1ms−1 and f = 0.1ms−1. The firing rate of each node is
about 50Hz. Parameters are chosen [9] as xex = 14/3, xin = −2/3, σ ex = 2ms, σ in = 5ms, τ = 20ms, xth = 1, xr = 0, and
τref = 2ms, Sex
i j = Sin
i j = 0.02.
large network (e.g., ∼ 100 nodes) are very slow, e.g., Monte
Carlo based methods [16, 21]. Our undergoing work is explor-
ing a fast algorithm that exploits the sparsity of effective in-
teractions. We have seen an indication that the algorithm can
work well for an I&F network with sparse coupling structure;
however, that work is yet to be fully verified to be conclusive.
ACKNOWLEDGMENTS
The authors thank David W. McLaughlin for helpful dis-
cussions. This work was supported by NSFC-11671259,
NSFC-11722107, NSFC-91630208 and Shanghai Rising-
Star Program-15QA1402600 (D.Z.); by NSF DMS-1009575
and NSFC-31571071 (D.C.); by Shanghai 14JC1403800,
15JC1400104, and SJTU-UM Collaborative Research Pro-
gram (D.C. and D.Z.); and by the NYU Abu Dhabi Institute
G1301 (Z.X., D.Z., and D.C.).
FIG. 3: Network of Erdos-Renyi connection. We randomly
generate 1000 networks of 100 nodes with an Erdos-Renyi
connection. The connection probability between two nodes is
0.05. The number of non-zero effective interaction v.s.
effective interaction order. The mean and standard deviation
are respectively shown by the black line and shaded area.
effective interactions. Correlated inputs can induce non-zero
high-order effective interactions [12]. It is yet to consider how
the statistics of inputs affect the sparsity of effective inter-
actions. Second, current algorithms for estimating non-zero
effective interactions (not limited to the second-order) for a
1234123410−410−2100 DepIndRDepRInd(34)(24)(23)(14)(13)(12)10−410−2100Interaction datashuffle(123)(124)(134)(234)(1234)10−410−2100Interaction datashuffle1234123410−410−2100 DepIndRDepRInd(34)(24)(23)(14)(13)(12)10−410−2100Interaction datashuffle(123)(124)(134)(234)(1234)10−410−2100Interaction datashuffle5
[1] S.-I. AMARI, Information geometry on hierarchy of probabil-
ity distributions, Information Theory, IEEE Transactions on, 47
(2001), pp. 1701 -- 1711.
[2] A. K. BARREIRO, J. GJORGJIEVA, F. RIEKE, AND E. SHEA-
BROWN, When do microcircuits produce beyond-pairwise cor-
relations?, Frontiers in computational neuroscience, 8 (2014).
[3] E. BULLMORE AND O. SPORNS, Complex brain networks:
graph theoretical analysis of structural and functional systems,
Nature Reviews Neuroscience, 10 (2009), pp. 186 -- 198.
[4] T. BURY, Statistical pairwise interaction model of stock mar-
ket, The European Physical Journal B, 86 (2013), p. 89.
[5] D. CAI, A. V. RANGAN, AND D. W. MCLAUGHLIN, Archi-
tectural and synaptic mechanisms underlying coherent spon-
taneous activity in v1, Proceedings of the National Academy
of Sciences of the United States of America, 102 (2005),
pp. 5868 -- 5873.
[6] Y. DAN, J.-M. ALONSO, W. M. USREY, AND R. C. REID,
Coding of visual information by precisely correlated spikes in
the lateral geniculate nucleus, Nature neuroscience, 1 (1998),
pp. 501 -- 507.
[7] E. GANMOR, R. SEGEV, AND E. SCHNEIDMAN, The architec-
ture of functional interaction networks in the retina, The journal
of neuroscience, 31 (2011), pp. 3044 -- 3054.
[8] E. GANMOR, R. SEGEV, AND E. SCHNEIDMAN, Sparse low-
order interaction network underlies a highly correlated and
learnable neural population code, Proceedings of the National
Academy of Sciences, 108 (2011), pp. 9679 -- 9684.
[9] W. GERSTNER AND W. M. KISTLER, Spiking neuron models:
Single neurons, populations, plasticity, Cambridge university
press, 2002.
[10] M. P. KARLSSON AND L. M. FRANK, Awake replay of re-
mote experiences in the hippocampus, Nature neuroscience, 12
(2009), p. 913.
[11] D. C. KNILL AND A. POUGET, The bayesian brain: the role
of uncertainty in neural coding and computation, TRENDS in
Neurosciences, 27 (2004), pp. 712 -- 719.
[12] J. H. MACKE, M. OPPER, AND M. BETHGE, Common in-
put explains higher-order correlations and entropy in a simple
model of neural population activity, Physical Review Letters,
106 (2011), p. 208102.
[13] O. MARRE, S. EL BOUSTANI, Y. FR ´EGNAC, AND A. DES-
TEXHE, Prediction of spatiotemporal patterns of neural activity
from pairwise correlations, Physical review letters, 102 (2009),
p. 138101.
[14] E. A. MARTIN, J. HLINKA, AND J. DAVIDSEN, Pairwise net-
work information and nonlinear correlations, Physical Review
E, 94 (2016), p. 040301.
[15] R. E. MIROLLO AND S. H. STROGATZ, Synchronization of
pulse-coupled biological oscillators, SIAM Journal on Applied
Mathematics, 50 (1990), pp. 1645 -- 1662.
[16] H. NASSER, O. MARRE, AND B. CESSAC, Spatio-temporal
spike train analysis for large scale networks using the maximum
entropy principle and monte carlo method, Journal of Statistical
Mechanics: Theory and Experiment, 2013 (2013), p. P03006.
[17] M. E. NEWMAN, The structure and function of complex net-
works, SIAM review, 45 (2003), pp. 167 -- 256.
[18] I. E. OHIORHENUAN, F. MECHLER, K. P. PURPURA, A. M.
SCHMID, Q. HU, AND J. D. VICTOR, Sparse coding and high-
order correlations in fine-scale cortical networks, Nature, 466
(2010), pp. 617 -- 621.
[19] E. SCHNEIDMAN, M. J. BERRY, R. SEGEV, AND W. BIALEK,
Weak pairwise correlations imply strongly correlated network
states in a neural population, Nature, 440 (2006), pp. 1007 --
1012.
[20] Y.
Y.
SHEMESH,
FORKOSH,
T. SHLAPOBERSKY, A. CHEN, AND E. SCHNEIDMAN,
High-order social interactions in groups of mice, Elife, 2
(2013), p. e00759.
SZTAINBERG,
O.
[21] J. SHLENS, G. D. FIELD, J. L. GAUTHIER, M. GRESCHNER,
A. SHER, A. M. LITKE, AND E. CHICHILNISKY, The struc-
ture of large-scale synchronized firing in primate retina, The
Journal of Neuroscience, 29 (2009), pp. 5022 -- 5031.
[22] J. SHLENS, G. D. FIELD,
J. L. GAUTHIER, M.
I.
GRIVICH, D. PETRUSCA, A. SHER, A. M. LITKE, AND
E. CHICHILNISKY, The structure of multi-neuron firing pat-
terns in primate retina, The Journal of neuroscience, 26 (2006),
pp. 8254 -- 8266.
[23] J. STRICKER, S. COOKSON, M. R. BENNETT, W. H.
MATHER, L. S. TSIMRING, AND J. HASTY, A fast, robust and
tunable synthetic gene oscillator, Nature, 456 (2008), pp. 516 --
519.
[24] A. TANG, D. JACKSON, J. HOBBS, W. CHEN, J. L. SMITH,
H. PATEL, A. PRIETO, D. PETRUSCA, M. I. GRIVICH,
A. SHER, ET AL., A maximum entropy model applied to spatial
and temporal correlations from cortical networks in vitro, The
Journal of Neuroscience, 28 (2008), pp. 505 -- 518.
[25] W. E. VINJE AND J. L. GALLANT, Sparse coding and decor-
relation in primary visual cortex during natural vision, Science,
287 (2000), pp. 1273 -- 1276.
[26] Z. WANG, Y. MA, F. CHENG, AND L. YANG, Review of pulse-
coupled neural networks, Image and Vision Computing, 28
(2010), pp. 5 -- 13.
[27] T. WATANABE, S. HIROSE, H. WADA, Y. IMAI, T. MACHIDA,
I. SHIROUZU, S. KONISHI, Y. MIYASHITA, AND N. MA-
SUDA, A pairwise maximum entropy model accurately de-
scribes resting-state human brain networks, Nature communi-
cations, 4 (2013), p. 1370.
[28] Z.-Q. J. XU, G. BI, D. ZHOU, AND D. CAI, A dynamical state
underlying the second order maximum entropy principle in neu-
ronal networks, Communications in Mathematical Sciences, 15
(2017), pp. 665 -- 692.
[29] D. ZHOU, Y. XIAO, Y. ZHANG, Z. XU, AND D. CAI, Causal
and structural connectivity of pulse-coupled nonlinear net-
works, Physical review letters, 111 (2013), p. 054102.
|
1204.0402 | 1 | 1204 | 2012-04-02T13:17:51 | Chemo-mechanical instabilities in polarizable active layers | [
"physics.bio-ph",
"nlin.PS"
] | We formulate and explore a generic continuum model of a polarizable active layer with neo-Hookean elasticity and chemo-mechanical interactions. Homogeneous solutions of the model equations exhibit a stationary long-wave instability when the medium is activated by expansion, and an oscillatory short-wave instability in the case of compressive activation. Both regimes are investigated analytically and numerically. The long-wave instability initiates a coarsening process, which provides a possible mechanism for the establishment of permanent polarization in spherical geometry. | physics.bio-ph | physics | Chemo-mechanical instabilities in polarizable active layers
M. H. Kopf1 and L. M. Pismen1, 2
1Department of Chemical Engineering, Technion -- Israel Institute of Technology, Haifa 32000, Israel
2Minerva Center for Nonlinear Physics of Complex Systems
We formulate and explore a generic continuum model of a polarizable active layer with neo-Hookean elas-
ticity and chemo-mechanical interactions. Homogeneous solutions of the model equations exhibit a stationary
long-wave instability when the medium is activated by expansion, and an oscillatory short-wave instability in the
case of compressive activation. Both regimes are investigated analytically and numerically. The long-wave in-
stability initiates a coarsening process, which provides a possible mechanism for the establishment of permanent
polarization in spherical geometry.
2
1
0
2
r
p
A
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
2
0
4
0
.
4
0
2
1
:
v
i
X
r
a
PACS numbers: 87.18.Hf, 82.40.Ck, 87.17.Aa, 83.10.Ff
Active media are capable to modify their mechanical prop-
erties under the influence of chemical signaling; in their turn,
chemical transformations may be affected by local stress,
strain and polarization. This creates a feedback loop that, on
the one hand, makes rheology of the medium flexible and re-
sponsive to the environment and, on the other hand, may cause
various instabilities, leading to spatial inhomogeneities, oscil-
lations, and spontaneous motion. The primary example of an
active biological medium is cytoskeleton, which is the princi-
pal structural element of eukaryotic cells responsible for their
integrity, reshaping, proliferation, and motion [1, 2].
Experimental studies of living tissues give many examples
of instabilities involving chemo-mechanical interactions, such
as spontaneous oscillatory contraction of muscle fibers [3],
traveling waves [4], and intermittent localized oscillations in
vivo [5] that occur, in particular, during such tissue reshap-
ing processes as dorsal closure [6, 7] and elongation of the
Drosophila egg shell [8].
Inhomogeneities and patterns of
non-genetic origin are inconspicuous in vivo but are promi-
nent in model systems, seen as clusters, swirls and intercon-
nected bands in motility assays [9], localized motor activity
spots in reconstituted actin-myosin mixtures [10], and veloc-
ity spurts in spreading tissues [11]. Mechanical impact on
chemical oscillations (though without feedback reaction) has
been also detected in inorganic gels [12].
Mechanisms of chemo-mechanical interactions in living
cells and tissues are variegated and largely unknown but un-
derlying general characteristics can be understood by taking
as a basis well-established principles of continuous mechan-
ics that account for the symmetries of the medium, and im-
posing a simple chemical feedback. The continuum approxi-
mation is particularly suitable for analytical studies of insta-
bility thresholds. Symmetry-breaking and oscillatory instabil-
ities due to chemo-mechanical coupling have been detected in
this way in the context of calcium transfer dynamics [13, 14],
and much later with myosin motors acting as chemical vari-
ables [15]. Further studies included nematic polarization in
the continuum theory [16, 17] based on de Gennes' descrip-
tion of nematic liquid crystals [18] and adding a term propor-
tional to chemical potential to account for the biological activ-
ity. The resulting equations of polar gels are quite formidable
but their simplified versions have enabled elegant description
of structures and defects in nematic gels [17], spontaneous
flow [19, 20], and other phenomena in polar active media.
In this Letter, we explore an alternative case of a medium
with vector polarization. The source of polarization of cy-
toskeleton is directionality of actin filaments characterized by
a unique direction of treadmilling and motion of myosin mo-
tors. In acto-myosin stress fibers, nematic order prevails due
to periodic polarity alternation as, for example, in sarcomeric
structures. On the other hand, stress fibers in motile cells have
a graded polarity pattern [21] implying vector polarization.
This kind of polarization is essential for generation of active
forces that play a primary role in tissue spreading [22 -- 24].
Another distinctive feature of our approach is allowing for
spontaneous emergence of polarization, alongside with me-
chanical deformation and chemical activity, whereas in earlier
studies of polar active media, with few exceptions [25], only
the direction, rather than the absolute value of the nematic
order parameter was evolving dynamically. After formulat-
ing dynamic equations applicable to a thin flat polarizable de-
formable layer on a solid substrate, we will determine analyt-
ically the various instabilities thresholds, and further explore
numerically the evolution of emerging patterns using the neo-
Hookean formulation applicable to strongly deformable soft
matter [26].
Dynamics of thin layers dominated by friction is Aris-
totelian (overdamped), with velocity rather than acceleration
proportional to the force. Close to equilibrium, the force driv-
ing the evolution of any dynamic variable Ψ can be derived
from an appropriate energy functional F{Ψ}, leading to a dy-
namic equation in the variational form Γ DtΨ = −δF/δΨ,
where Dt = ∂t+v·∇ is the substantial derivative, v is the local
velocity, and Γ is an appropriate friction coefficient. We will
write in this form basic dynamic equations for the deforma-
tion and polarization fields, and further add non-equilibrium
terms accounting for the active character of the medium.
The medium is assumed to be incompressible but it can
stretch in the horizontal plane x, while modifying the layer
thickness h. Since the vertical deformation is small compared
to the deformations in the horizontal plane and the latter, in
turn, are uniform in the vertical direction, the elastic energy is
expressed through the two-dimensional (2D) strain tensor. Let
x, X(x) be the dimensionless position vectors, respectively,
(cid:18) ωjk
(cid:19)
in the reference (stress-free) and deformed states. The com-
ponents of the strain tensor are ωij = ∂iXk∂jXk − δij, where
δij is the Kronecker delta; summation over repeated indices is
presumed here and below. The elastic energy is
(cid:90)
Fel =
Leld2x, Lel = µhL0, L0 =
1
4
ωijωij.
(1)
where µ is the shear modulus.
Unlike the standard 2D formulation, the variation of h
should be taken into account when the dynamic equations are
derived. In view of the incompressibility, h = h0J−1, where
h0 is the unperturbed thickness and J is the determinant of the
2D matrix with the components ∂jXi. We write the dynamic
equations in the dimensionless form
η DtXi =
1
2
∂j
∂kXi − L0
J 2 ikjl∂lXk
J
+
fi
2J
,
(2)
where F = (µ/ax) f is an active force per unit volume (to be
x/(2µh0at), ax, at are the length
specified below), η = Γela2
and time scales, and ik is the 2D antisymmetric matrix.
integral over the deformed domain: Fp = (cid:82)
2 p2(cid:17)
Planar polarization is characterized by the 2D vector order
parameter p(X, t). The polarization energy is expressed as an
Lpd2X, where
the Lagrangian Lp should be expressed through rotationally
invariant combinations of p and the 2D gradient operator ∇
with respect to the coordinates Xi. Retaining terms with the
added orders in p and ∇ not exceeding four, we write
+ 2κ1∇ipi
+ (∇ipi)2 + κ2(ik∇ipk)2 + κ3p2∇ipi
(cid:104)
α2p2(cid:16)
Lp =
κh
2a2
x
(cid:3) .
1 +
(3)
κ0
−1/2
0
The second term is irrelevant, unless the coefficient κ1 cou-
ples the order parameter with another field.
In an active
medium, this might be the concentration c of some chemi-
−1/2
cal species. We choose κ
as the polarization scale and set
0
c, where c(X, t) is interpreted as a scaled devi-
κ1 = −κ
ation from a reference concentration c0. Setting also κ3 = 0,
we write the dynamic equation generated by the Lagrangian
(3) in the dimensionless form
−1 (∇jpi − ∇ipj)(cid:3)
(cid:2)J
−1(cid:0)1 + p2(cid:1) pi,
−1∇jpj) + κ2∇j
γ Dtpi = ∇i(J
−1c) − α2J
− ∇i(J
where γ = Γpa2
x/(κh0at). Generally, one has to include
the variation of the polarization energy into the deformation
equations; indeed, a polarized medium may deform to releave
polarization gradients. We will avoid, however, this compli-
cation, assuming κ (cid:28) µ. The coupling between polarization
and deformation will be established instead by the active force
in Eq. (2) directed along the polarization direction, fi = 2qpi.
The feedback loop is closed by making concentration of
the active species strain-dependent. We write the equation of
the concentration field taking into account that, due to incom-
pressibility, the concentration is not affected by deformation;
(4)
2
FIG. 1: Stability diagram in the parametric plane α, βq. The oscil-
latory instability limits are shown for γ = 5/4 and η = 1/3 (solid
line), η = 5/3 (dashes), and η = 10/3 (dots). Inset: the critical
wavelength k at the oscillatory bifurcation lines parameterized by α.
the diffusional flux is, however, proportional to the local thick-
ness. Allowing for the production rate proportional to the lo-
cal strain and linear decay with the constant identified with the
inverse time scale 1/at, we write the concentration equation
in the dimensionless form
Dtc = J∇i(J
−1∇ic) − c + β(J − 1).
(5)
where we have chosen as the length scale the diffusion length
ax = √Dat, D being the diffusivity.
The above equation system becomes extremely convoluted
if applied in a fixed coordinate frame where, besides com-
puting the substantial derivative that includes the velocity
v = ∂tX, one would need to transform the ∇ operator in
Eqs. (4), (5) to account for deformations. These complications
do not affect, however, linear stability analysis, and will be
avoided at large deformations by implementing a Lagrangean
numerical procedure.
We explore now stability of the homogeneous solution p =
0, c = 0, X = x to infinitesimal perturbations. The linearized
system determining the bifurcation threshold can be expressed
through the 2D strain ψ = J − 1 = ∇i(Xi − xi) and splay
φ = ∇ipi by taking the divergence of linearized Eqs. (2), (4):
(6)
(7)
(8)
η ∂tψ = ∇2ψ + qφ,
γ ∂tφ = ∇2φ − α2φ − ∇2c,
∂tc = ∇2c − c + βψ.
A nontrivial solution of Eqs. (6) -- (8) is sought for as a wave-
form exp(λt + i k · x) with a wave vector k. The eigenvalues
λ determining stability of the uniform state c = ψ = φ = 0
are computed by setting to zero the determinant of the Fourier
transformed linearized system. The bifurcation thresholds are
computed in a standard way by requiring supk((cid:60)λ) = 0.
In the case of extentional activation (β > 0), one can show
that the branch of real eigenvalues first reaches zero in the
oscillatoryinstabilitystationaryinstabilitystable-200-150-100-500500123456βqα00.40.81.201234kα3
FIG. 2: Snapshots from a simulation in the oscillatory unstable para-
metric region (βq = −210, (βq)c ≈ −196). The time difference
√
√
between the two images equals half the oscillation period. The sim-
5 square. Lighter
ulation domain is a double-periodic 10
shades correspond to higher concentration levels; scaled arrows indi-
cate the polarization p.
5 × 10
FIG. 3: Histogram of the angle ϕ between the directions of the
elastic and active force.
long-wave mode k = 0 at βq = α2.
In the case of com-
pressive activation (β < 0), a wave instability at finite k and
(cid:61)λ (cid:54)= 0 is observed. The analytic formulas for the critical
wavenumber kc and the critical value βc are too cumbersome
to be of practical use and will not be listed here. The min-
imal gap width between the two instability limits equals to
γ + η. The bifurcation diagram in the parametric plane α, βq
is shown in Fig. 1.
We explore dynamics beyond the bifurcation thresholds for
both types of activation by direct numerical simulation, start-
ing from randomly perturbed homogeneous initial conditions.
We discretize Eqs. (2), (4), and (5) using a flexible grid with
nodes following the physical displacement of the medium. In
the spirit of mesh-free computations [27], the spatial deriva-
tives are calculated based on local approximation of the fields
by low-order polynomials. The resulting set of ordinary dif-
ferential equations is then solved using an embedded adaptive
Runge-Kutta-Fehlberg scheme of the order 4 and 5 [28]. The
parameters are set to q = √5/100, α = √5, κ2 = 1, η =
5/3, γ = 5/4.
In the oscillatory regime, we observe a wave pattern with
the active force (directed from high to low c regions) follow-
ing the rotating concentration gradient between the domains
oscillating in antiphase, as seen in Fig. 2. The directions of
the elastic and active force are, however, disaligned due to de-
lay, so that the angle between the two is widely distributed
with a shallow maximum about π/2 (Fig. 3).
In the case of extensional activation, the behavior is qual-
itatively different. Initially, a short-wave pattern correspond-
ing to a fastest-growing mode emerges from the initial noise.
Then, a slow coarsening process ensues, during which larger
domains grow while smaller ones either merge or vanish, so
that the dominant wavelength grows, as seen in Fig. 4. The
wavelength corresponding to the maximum of the Fourier
spectrum is plotted against time in Fig. 5. Due to the finite
size of the simulation domain, the plot for each individual run
has a number of discontinuous steps but it approaches a mono-
tonically growing curve when averaged over 30 runs.
FIG. 4: Domain coarsening: snapshots at t = 4 × 103 (a), 2 ×
104 (b), 3 × 104 (c), 5 × 104 (d), simulation on the double-periodic
√
√
5 × 100
100
5 square domain for βq = 6, (βq)c = 5.
FIG. 5: Time evolution of the wavelength of the dominant Fourier
√
√
mode during coarsening (the average of 30 simulations on the
5 × 100
double-periodic 100
5 square domain for βq = 6).
00.511.522.53ϕ5010015020025001000020000300004000050000λtbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbbb4
[1] J. Howard, Mechanics of Motor Proteins and the Cytoskeleton
(Sinauer Associates, Sunderland, Mass., 2001).
[2] O. Lieleg, M. M. A. E. Claessens, and A. R. Bausch, Soft Mat-
ter 6, 218 (2010).
[3] H. Fujita and S. Ishiwata, Biophys. J. 75, 1439 (1998).
[4] G. Gerisch, T. Bretschneider, and A. Muller-Taubenberger, Bio-
phys. J. 87, 3493 (2004).
[5] A. C. Martin, Developmental Biol. 341, 114 (2010).
[6] J. Solon, A. Kaya-C¸ opur, J. Colombelli, and D. Brunner, Cell
[7] N. Gorfinkiel, S. Schamberg, and G. B. Blanchard, Genesis 49,
[8] L. He, X. Wang, H. L. Tang, and D. J. Montell, Nature Cell
137, 1331 (2009).
522 (2011).
Biol. 12, 1133 (2010).
Nature 467, 73 (2010).
[9] V. Schaller, C. Weber, C. Semmrich, E. Frey, and A. R. Bausch,
[10] M. S. e. Silva, M. Depken, B. Stuhrmann, M. Korsten, F. C.
MacKintosh, and G. H. Koenderink, Proc. Natl Acad. Sci.
U.S.A. 108, 9408 (2011).
[11] L. Petitjean, M. Reffay, E. Grasland-Mongrain, M. Poujade,
B. Ladoux, A. Buguin, and P. Silberzan, Biophys J 98, 1790
(2010).
[12] O. Kuksenok, V. V. Yashin, and A. C. Balazs, Soft Matter 3,
1138 (2007).
Biol. 1, 51 (1984).
[13] J. D. Murray and G. F. Oster, IMA J. Math. Appl. Math. Med.
[14] G. Oster and G. Odell, Physica D 12, 333 (1984).
[15] S. Banerjee and M. C. Marchetti, Soft Matter 7, 463 (2011).
[16] K. Kruse, J. F. Joanny, F. Julicher, J. Prost, and K. Sekimoto,
Eur. Phys. J. E 16, 5 (2005).
[17] F. Julicher, K. Kruse, J. Prost, and J.-F. Joanny, Phys. Rep. 449,
[18] P. de Gennes and J. Prost, The Physics of Liquid Crystals (Ox-
ford University Press, Oxford, U.K., 1993).
[19] K. Doubrovinski and K. Kruse, Phys. Rev. Lett. 107, 258103
3 (2007).
(2011).
[20] S. Furthauer, M. Neef, S. W. Grill, K. Kruse, and F. Julicher,
New J. Phys. 14, 023001 (2012).
[21] S. Deguchi and M. Sato, Biorheology 46, 93 (2009).
[22] D. L. Nikolic, A. N. Boettiger, D. Bar-Sagi, J. D. Carbeck, and
S. Y. Shvartsman, 291, C68 (2006).
[23] M. Poujade, E. Grasland-Mongrain, A. Hertzog, J. Jouanneau,
P. Chavrier, B. Ladoux, A. Buguin, and P. Silberzan, Proc. Nat.
Acad. Sci. USA 104, 15988 (2007).
[24] M. Salm and L. M. Pismen, Phys. Biol. 9 (2012).
[25] G. Salbreux, J. Prost, and J. F. Joanny, Phys. Rev. Lett. 103,
058102 (2009).
[26] R. W. Ogden, in Nonlinear Elasticity: Theory and Applications,
edited by Y. B. Fu and R. W. Ogden (Cambridge University
Press, Cambridge, U.K., 2001), pp. 1 -- 58.
[27] E. Onate, S. Idelsohn, O. C. Zienkiewicz, and R. L. Taylor, Int.
J. Numer. Meth. Engng. 39, 3839 (1996).
[28] W. H. Press, Numerical Recipes in C (Cambridge University
Press, Cambridge, U.K., 1999).
[29] C. R. Cowan and A. A. Hyman, Development 134 (2007).
[30] N. W. Goehring, P. K. Trong, J. S. Bois, D. Chowdhury, E. M.
Nicola, A. A. Hyman, and S. W. Grill, Science 334, 1137
(2011).
FIG. 6: Cross section through the dashed line in Fig. 4 (d). The
polarization component along the line p · ey and deformation J − 1
are scaled to show all three fields in the same plot.
√
FIG. 7: Time evolution of the concentration c at different polar angles
θ on a sphere of radius 10
5.
Unlike the oscillatory case, the active and elastic forces are
almost perfectly anti-aligned: the mutual angle deviates from
π by less than 3.5◦ with 98% probability. As the active force
is directed against the concentration gradient, the low-c re-
gions are squeezed into narrow stripes separating high-c do-
mains.
In turn, the chemical production is inhibited within
these stripes due to strong compression, as the deformation
acts as the source of c in Eq. (5). The resulting correlation
between J and c is seen in Fig. 6 that shows the cross-section
profiles at a late stage of coarsening.
The coarsening process provides a natural mechanism for
polarity establishment in cortical tissues, as is observed, for
example, in the C. elegans embryo [29]. When Eqs. (2),
(4), and (5) are solved on a spherical surface, the coarsen-
ing consolidates two domains, whereafter the smaller one is
compressed into a low-c spot. This is illustrated in Fig. 7,
where the evolution of c is plotted for five different polar an-
gles θ, where θ = 0 is defined by the location of the high-
concentration pole. As the polarization is aligned with the
concentration gradient, the active force is directed from one
pole to the other. The established polarity axis is permanent,
unlike Ref. [30] where advection serves as a mere trigger re-
quiring additional means for preserving the polarization.
This work has been supported by the Human Frontier Sci-
ence Program (Grant RGP0052/2009-C).
-60-40-2002040050100150200yc50(ey·p)100(J−1)-200-10001002003004005006007008000200040006000800010000ctθ=0θ=π/4θ=π/2θ=3π/4θ=π |
1212.1673 | 2 | 1212 | 2013-11-14T11:43:39 | Parasites on parasites: coupled fluctuations in stacked contact processes | [
"physics.bio-ph",
"cond-mat.stat-mech",
"q-bio.PE"
] | We present a model for host-parasite dynamics which incorporates both vertical and horizontal transmission as well as spatial structure. Our model consists of stacked contact processes (CP), where the dynamics of the host is a simple CP on a lattice while the dynamics of the parasite is a secondary CP which sits on top of the host-occupied sites. In the simplest case, where infection does not incur any cost, we uncover a novel effect: a nonmonotonic dependence of parasite prevalence on host turnover. Inspired by natural examples of hyperparasitism, we extend our model to multiple levels of parasites and identify a transition between the maintenance of a finite and infinite number of levels, which we conjecture is connected to a roughening transition in models of surface-growth. | physics.bio-ph | physics | epl draft
3
1
0
2
v
o
N
4
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
3
7
6
1
.
2
1
2
1
:
v
i
X
r
a
Parasites on parasites: coupled fluctuations in stacked contact
processes
Steven J. Court, Richard A. Blythe and Rosalind J. Allen
SUPA, School of Physics and Astronomy, University of Edinburgh, Mayfield Road, Edinburgh EH9 3JZ, UK
PACS 02.50.Ey -- Stochastic processes
PACS 05.70.Fh -- Phase transitions: general studies
PACS 87.23.Cc -- Population dynamics and ecological pattern formation
Abstract -- We present a model for host-parasite dynamics which incorporates both vertical
and horizontal transmission as well as spatial structure. Our model consists of stacked contact
processes (CP), where the dynamics of the host is a simple CP on a lattice while the dynamics of
the parasite is a secondary CP which sits on top of the host-occupied sites. In the simplest case,
where infection does not incur any cost, we uncover a novel effect: a nonmonotonic dependence of
parasite prevalence on host turnover. Inspired by natural examples of hyperparasitism, we extend
our model to multiple levels of parasites and identify a transition between the maintenance of a
finite and infinite number of levels, which we conjecture is connected to a roughening transition
in models of surface-growth.
Introduction. -- The need to understand and control
the dynamics of infections has motivated the development
of a variety of statistical mechanical models. One of the
most important of these is the contact process (CP) [1 -- 4],
which describes the dynamics of an infection in a spatially
structured population. In the contact process, each site
on a lattice represents a host organism which can be in-
fected or susceptible (uninfected). Infection is transmitted
to neighbouring susceptible host sites at rate b, and in-
fected hosts recover (i.e. become susceptible) at rate d.
The contact process provides a baseline model for many
problems in ecology and epidemiology. It is also of funda-
mental importance in statistical physics. This is because
it exhibits a non-equilibrium phase transition between an
infected and a non-infected phase at a critical value of
λ = b/d [2, 3].
The standard CP model for an infected population as-
sumes that the host population is of fixed size, without
turnover, i.e. host births and deaths. This is valid if the
timescale on which the infection is gained and lost is much
shorter than the lifespan of an individual. Some infections,
however, are carried by individuals for long times, and may
be transmitted to offspring upon reproduction (vertical
transmission), as well as being transmitted horizontally --
i.e. upon physical contact between individuals. Examples
include plasmids carried by bacteria [5], some microsporid-
ian parasites of insects [6] and pathogens including HIV
and several hepatitis viruses. Here, we investigate the
interplay between vertical and horizontal transmission in
such populations. Other authors have determined the con-
ditions for parasite persistence in mean-field models that
lack spatial structure, in which the parasite may affect
host fitness [7 -- 9], and have suggested that these condi-
tions may be affected by spatial structure [10]. Here, we
show that spatial structure can produce a qualitatively
new effect: a coupling between the dynamics of the infec-
tion and of the underlying host population, even when the
infection does not affect the fitness of the host.
We present a two level stacked contact process, in which
the host population is represented by a CP on a lattice,
and the parasite population is represented by a second
CP which sits on top of the host CP. This model in-
corporates both vertical and horizontal transmission --
if an infected host reproduces or dies, the parasite is re-
produced or dies with it (vertical transmission), and a
parasite-infected site can infect a neighbouring site if it is
occupied by an uninfected host (horizontal transmission).
We characterize the conditions for parasite persistence in
the form of a phase diagram and discover an interesting
phenomenon: although the steady-state properties of the
host population depend only on the ratio of its birth and
death rates, λ = b/d, the prevalence of the parasite can de-
pend non-monotonically on the host population's turnover
rate. This phenomenon has its origins in the fluctuations
p-1
S. J. Court et al.
of spatial clusters of host individuals, and cannot straight-
forwardly be captured by a mean-field theory.
Parasitic infections are not always limited to two levels.
Hyperparasitism, in which an organism carrying a primary
parasite is susceptible to a secondary parasite [11], can be
harnessed as a biocontrol mechanism -- examples include
viral infections of the fungus Cryphonectria parasitica that
causes chestnut blight [12], and cytoplasmic RNA elements
that infect the fungus causing Dutch elm disease [13]. In-
spired by these scenarios, we extend our model to a multi-
level stacked contact process, in which individuals carrying
a primary infection may be susceptible to secondary infec-
tions, and those carrying the secondary infection may be
susceptible to tertiary infections, etc. This raises a number
of questions: How does the dynamics of one parasite level
couple to the next, and how many levels of parasites are
sustainable in a population? We show that in our stacked
contact process, there is a well-defined transition between
maintenance of an infinite hierarchy of levels of parasites,
and limitation to a finite number of levels. The transition
between these two regimes appears to be connected to a
roughening transition in certain surface growth processes
[14 -- 16]. This work presents new challenges to our under-
standing of contact processes, with potential implications
for the dynamics of long-timescale infections.
Two-level stacked contact process. -- We begin
by considering the dynamics of a host and a single para-
site which can be transmitted vertically or horizontally --
a two-level stacked CP -- on a 2D square lattice. Lattice
sites can be either empty or occupied by a host organism.
Occupied lattice sites become empty at rate d0, due to
death of hosts, and host organisms attempt to replicate
into neighbouring empty sites at rate b0 (the attempt rate
in any direction is b0/z, z being the coordination number
of the lattice). Host organisms may also carry a parasite,
which we assume to be neutral in the sense that organ-
isms with and without the parasite have identical birth
and death rates. Infected individuals pass the parasite to
uninfected neighbours at rate b1, and the parasite is lost
at a constant rate d1. We assume that the offspring of an
infected organism is also infected (vertical transmission),
and that when an infected host dies, the parasite dies with
it. Thus the infection dynamics consists of a secondary
CP, with parameters b1 and d1, which sits on top of the
host's birth-death CP. Since the host population does not
occupy the whole lattice, the secondary CP does not take
place on a regular lattice but rather on the irregular net-
work of occupied sites which changes stochastically due to
the dynamics of the host population. The processes that
occur in our model are illustrated in Figure 1. We sim-
ulate this model using a stochastic kinetic Monte Carlo
algorithm, in which all events occur as Poisson processes
[17], on a 100 × 100 lattice with periodic boundary condi-
tions. Averages are calculated after an initial transient to
allow the steady state to be reached.
b0
z
b0
z
b1
z
d0
d0
d1
Figure 1: The events that constitute the 2-level stacked
CP. Red and black circles represent host organisms and
parasites respectively. A site can be empty, occupied by
a susceptible host or occupied by an infected host. The
left panels represent reproduction of a susceptible or in-
fected host (top and middle, rate b0), and transmission of
the parasite (bottom, rate b1). The right panels represent
death of a susceptible or infected host (top and middle,
rate d0), and loss of the parasite (bottom, rate d1). z is
the number of nearest neighbour sites. The dotted squares
indicate the level of the CP at which the event happens.
Host dynamics influences parasite persistence. --
In this model, the host population undergoes a simple CP,
whose stationary properties are fully determined by the
single parameter λ0 = b0/d0 [1 -- 4]. If λ0 is less than a crit-
ical value, λcrit, the only steady state is an empty lattice
and the population rapidly becomes extinct. If λ0 > λcrit,
a non-zero population can be maintained for long times,
before eventually becoming extinct due to a rare fluctua-
tion. The transition between these two regimes is second
order; the values of λcrit and the associated critical ex-
ponents have been characterised in detail [3]. Defining
τi = 0 or 1 if site i is empty or occupied respectively, the
dynamics of the host population obeys
d
dt
hτii =
1
z X
j
b0 hτj (1 − τi)i − d0 hτii ,
(1)
where the sum is over the z neighbours of lattice site
i and the angle brackets denote averages over multiple
realizations of the dynamics. The steady-state solution
of Eq. (1) satisfies hτii = (λ0/(λ0 − 1)) hτiτji.
Ignor-
ing correlations between neighbouring sites by assuming
that hτiτji ≈ hτii hτji, we arrive at the mean-field result
hτ i = 1 − (1/λ0).
Considering now the dynamics of the parasite, we de-
fine σi = 1 if site i contains an infected host and σi = 0
otherwise (i.e. if site i is either unoccupied or contains a
susceptible host). σi obeys
d
dt
hσii =
[b0 hσj (1 − τi)i + b1 hσj (τi − σi)i]
1
z X
− d0 hσii − d1 hσii .
j
(2)
The first term on the r.h.s. of Eq. (2) corresponds to ver-
tical transmission: an empty site is filled by replication of
p-2
1
d
/
1
b
=
1
λ
t
c
n
i
t
x
e
t
s
o
H
12
8
4
0
Parasite persists
2
Parasite lost
6
10
λ
0 = b0 / d0
14
Figure 2: Phase diagram of the two-level stacked CP. Sym-
bols: simulation data showing boundary of parasite persis-
tence, for d1 = 1 and d0 = 1 (circles) or d0 = 20 (squares).
Dashed line: mean-field prediction λ1 = λ0/(λ0 − 1).
an infected host. The second term represents horizontal
transmission: a susceptible host, denoted by (τi − σi), is
infected by a parasite-carrying neighbour. The final two
terms correspond to death of an infected host and loss of
the parasite. Eq. (2) can be rewritten as
d
dt
hσii =
[b0 hσji + (b1 − b0) hτiσji − b1 hσiσji]
1
z X
− (d1 + d0) hσii .
j
(3)
Comparing this with Eq. (1), we see that the cross-
correlation hτiσj i perturbs the parasite dynamics from
that of a standard CP. Interestingly, however, if b1 = b0
(i.e.
if the rates of horizontal and vertical transmission
are equal), this term vanishes and the form of Eq. (3)
becomes that of a standard CP with parameter λeff =
λ0λ1/(λ0 + λ1) = b/ Pk dk. The parasite dynamics also
has CP-like behaviour in the mean-field limit, where we
neglect spatial correlations: setting hτiσj i ≈ hτii hσj i and
hσiσji ≈ hσii hσj i in Eq. (3), we obtain the steady-state
solution hσi = 1 − (1/λ0) − (1/λ1) ≡ 1 − (1/λeff ).
Figure 2 shows the phase diagram for our model, as a
function of λ0 and λ1. Three steady state scenarios are
possible: (i) the host population is extinct, (ii) the host
population is finite but the parasite is extinct, and (iii) the
parasite persists within a finite host population. Since the
dynamics of the host population is a standard CP, the host
population is extinct (scenario (i)) if λ0 < λcrit. For λ0 >
λcrit, the host population persists, with or without the
parasite. The condition for parasite persistence (i.e. the
boundary separating scenarios (ii) and (iii)) is predicted
by the mean field theory to be λ1 = λ0/(λ0 − 1) (obtained
by setting hσi = 0) -- shown by the dashed line in Figure
2. As one might expect, as λ0 decreases, the density of the
host population decreases, and a higher rate of horizontal
transmission (λ1) is needed to maintain the parasite.
The symbols in Figure 2 show the boundaries between
parasite persistence and loss obtained from our kinetic
Parasites on parasites
Monte Carlo simulations, for two values of d0, the host
death rate. The fact that these are shifted upwards and
to the right of the mean-field prediction shows that spatial
correlations make it harder to maintain the parasite. It is
also interesting that we obtain different results for the two
different values of d0. This shows that the parameters λ0
and λ1 do not fully determine the phase behaviour of the
system. For fixed λ0 = b0/d0, a higher turnover rate of the
host population (i.e. higher d0 and b0) apparently makes
it harder to maintain the parasite.
In fact the situation is more complex. Figure 3a shows a
more comprehensive investigation of the steady-state par-
asite density hσi as a function of the host turnover rate. In
these simulations, we vary b0 and d0 keeping λ0 = b0/d0
fixed. The steady-state density of the host population is
constant, since it depends only on λ0. Remarkably, for
some parameter combinations, the density of the parasite
actually depends non-monotonically on the host turnover
rate. If the host dynamics is slow, increasing the turnover
rate increases the parasite density, while if the host dy-
namics is fast, the parasite density decreases with the
turnover rate.
This dependence of the parasite density on the host
turnover rate is absent in the mean-field theory and must
therefore be a consequence of spatial correlations. To
investigate this, we measure in our simulations the spa-
tial correlations of the parasite density via the function
Ci,j = [hσiσji/hσi] − hσi. Here, the first term is the con-
ditional probability of finding a parasite at site j given
that there is one at site i, while the second term ensures
that Ci,j → 0 as i − j → ∞ (where i − j denotes the
distance between sites i and j). We then fit our data to
the functional form Ci,j ∼ e−i−j/ξ to extract the spa-
tial correlation length ξ, which can be thought of as the
parasite cluster size. This is shown in Figure 3b. Strik-
ingly, the correlation length is minimal at the turnover rate
where the parasite density is maximal. This makes intu-
itive sense: since horizontal transmission requires contact
0.4
<σ>
0.1
0
0.12
<σ>
0.08
2
ξ
1.8
1
1.5
0
0.5
d0
(a)
2
3
1
d0
(b)
Figure 3: (a) Parasite density hσi as a function of host
turnover rate d0 for fixed λ0 (and hence fixed host den-
sity). Curves top to bottom correspond to (λ0, b1, d1) =
(5, 4, 1),(2.5, 15, 1),(2, 8, 1) and (1.8, 15, 1).
(b) Parasite
density hσi (triangles, left axis) and spatial correlation
length ξ (squares, right axis) for (λ0, b1, d1) = (1.8, 15, 1).
p-3
S. J. Court et al.
between infected and uninfected hosts, minimal spatial
clustering of the parasite population results in maximal
parasite density.
Why should the turnover rate of the host affect the spa-
tial clustering of the parasite? At fast host turnover rates,
we observe in our simulations that local patches of host
organisms rapidly grow from a single seed and vanish by
stochastic extinction, on a faster timescale than that of
horizontal transmission. Because offspring are always of
the same type as their parents, organisms in a single patch
are either all infected or all uninfected. Thus the infected
population is highly clustered, leading to low rates of hori-
zontal transmission and low parasite density. By contrast,
when the host population turnover is slow, the parasite
dynamics constitutes a CP on the effectively frozen, dis-
ordered network of lattice sites that are occupied by host
organisms. This network contains clusters of sites that
may be disconnected, or poorly connected, from the rest
of the host population; this creates the possibility of local,
stochastic parasite extinctions. This is most obviously the
case in the low λ0 regime where the host density is low,
but it is still true for higher host densities. In this regime,
an increase in the host turnover rate serves as a mixing
mechanism, by more homogeneously distributing the host
population and thereby providing greater opportunity for
the parasite to spread.
Thus the nonmonotonic dependence of the parasite den-
sity on the host turnover rate arises from a coupling of the
spatial fluctuations of the underlying host CP to the par-
asite dynamics, and can be viewed as a competition be-
tween the mixing effect of the host birth-death process at
low turnover rates and the population segregation arising
from parent-offspring clustering at high turnover rates.
Multi-level stacked contact processes. -- Moti-
vated by natural examples of hyperparasitism, we now in-
vestigate a multi-level stacked CP, where individuals car-
rying a primary infection are susceptible to a secondary
infection, and those with the secondary infection are sus-
ceptible to tertiary infections, and so on. In the M -level
stacked CP, a site is labelled 0 if it is empty, 1 if it contains
an uninfected host, and m = 2, 3, . . . , M if it contains a
host infected with m − 1 levels of parasites. For example,
a site labelled 2 contains a host infected with only a pri-
mary infection (parasite) while a site labelled 3 contains
a host which is infected with primary and secondary in-
fections (a parasite and a hyperparasite). Note that the
presence of an infection at level m implies the presence
of lower-level infections (e.g. one cannot have a secondary
infection without a primary infection). Figure 4 illustrates
the dynamical processes that can occur in this model. An
empty site may be occupied at rate b0/z by reproduction of
a neighbouring site of any label (top left), or a site labelled
m may be infected at rate bm/z by higher-level parasites
from a neighbouring site with label n > m; its label is
then promoted to n (middle and bottom left). The host
organism of a level m site can die, at rate d0 (top right),
b0
z
b1
z
2b
z
d0
d1
d2
Figure 4: Possible transitions for the multi-level CP, con-
sidering a site with label m = 3. The symbols are as
in Fig. 1; the green circles represent the secondary para-
site. Birth processes are shown on the left. The organism
in question can reproduce into a neighbouring empty site
(top left), transmit its primary parasite to a susceptible
neighbour (middle left - note the secondary parasite is
also transmitted), or transmit its secondary parasite to a
susceptible neighbour (bottom left). Death processes are
shown on the right. These consist of death of the host
(top right), loss of the primary parasite (middle right) or
loss of the secondary parasite (bottom right).
or experience loss of one of its parasites, at rate dn (where
n ≤ m relates to the level of parasite that is lost). When
the latter happens, all higher parasites are also lost and
the site is demoted to level n (middle and bottom right
panels in Fig. 4). The parameters of the model are the
number of levels M , and the set of level-dependent birth
and death rates {b0, . . . , bM−1; d0, . . . , dM−1}.
We first consider the special case where the birth rates
at all levels are equal; bn = b ∀ n. We divide lattice sites
into two sets: sites that have labels greater than or equal
to m (here denoted M+), and sites that have labels less
than m (here denoted M−). An M− site can become
M+ by infection by a neighbouring M+ site: this occurs
at rate b/z regardless of the level of the M+ site. An M+
site can become M− by death of the host or by loss of a
parasite at levels n = 2 . . . m. This occurs at total rate d =
Pm−1
ℓ=0 dℓ where the sum is over levels up to m − 1. Thus
the dynamics of the M+ density is exactly equivalent to
that of a standard CP with parameter λeff = b/ Pm−1
ℓ=0 dℓ.
This observation provides us with important insight into
the number of levels that are sustainable in a stacked CP.
For the standard (one-level) CP, a finite density of oc-
cupied lattice sites can only be sustained for λ > λcrit,
where λcrit ≈ 1.649 for a 2D square lattice [3].
In the
stacked CP with equal birth rates we therefore expect
the density of M+ sites, those with label n ≥ m, to be
non-zero only if λeff = b0/ Pm−1
ℓ=0 dℓ < λcrit. Figure 5
shows simulation results for the average density of lat-
tice sites with label n ≥ m, as a function of m, for a
stacked CP with equal birth rates bn = b and death rates
dn = d = 1, for several values of b. By mapping onto a
standard CP with λeff = b/(md), we predict that all lev-
els n ≥ m∗ = ⌈b/(λcritd)⌉ have zero density and are thus
p-4
1
0.8
0.6
0.4
0.2
y
t
i
s
n
e
D
Parasites on parasites
0.6
0.4
y
t
i
s
n
e
D
0.2
f
=
1
f = 0.5
f = 0.6
f
=
0
.
7
5
l
e
v
e
l
e
t
i
n
i
f
n
i
f
o
y
t
i
s
n
e
D
0.6
0.2
0
0
2
6
4
Level m
8
10
2
4
6
Level m
8
(a)
0.2
0.6
d0
(b)
1
Figure 5: Density of sites with label m or higher, as a
function of level m, for a stacked CP with equal birth rates
bn = b and death rates dn = 1. Left to right correspond to
simulations with b = 2,4,6,10 and 15 respectively. Arrows
indicate m∗ = ⌈b/(λcritd)⌉, the boundary for sustainability
as predicted by mapping to the standard CP.
unsustainable. Our simulation results, shown in Figure
5, bear this out: the system indeed only sustains a finite
number m∗ of parasite levels.
Are there any circumstances where a stacked CP can be
sustained for an infinite number of levels? This is indeed
possible if either the birth rate bn increases, or the death
rate dn decreases, sufficiently strongly with n. We first
suppose that the birth rate bn = b is constant but the
death rate decreases by a factor f at successive levels:
dn = f dn−1 = f nd0 where 0 < f < 1.
In this case,
the average density of sites with label n ≥ m is given by
that of a standard CP with λeff = b/ Pm−1
ℓ=0 dℓ = b(1 −
f )/ [d0(1 − f m)]. To sustain an infinite number of levels,
we require that λm→∞
= b(1−f )/d0 ≥ λcrit, which implies
that f ≤ 1 − (d0λcrit/b). Our simulation results, Fig. 6a,
show that, with d0 = 1 and b = 5, the system indeed
sustains an infinite number of levels for values of f above
the critical value of f ≈ 0.67.
eff
Interestingly, this analysis also shows that, for given
values of b and f , there exists a critical value of the host
death rate d0 which separates regimes where the stacked
CP can and cannot sustain an infinite number of levels.
This is given by d∗
0 = b(1 − f )/λcrit. Figure 6b shows,
as a function of d0, the predicted density of sites with
label n ≥ m, as m → ∞: this is given by the density of
a standard CP with λeff = b(1 − f )/d0. The density as
m → ∞ indeed falls to zero at d0 = d∗
0.
Finally, we explore briefly the case where, instead of
varying the death rate between levels, we instead increase
the birth rate by a factor 1/f (i.e.
set bn = bn−1/f ).
In this case the dynamics no longer maps onto that of
a standard CP -- but nevertheless, our simulations show
qualitatively similar results (dashed lines in Figure 6a).
The transition to an infinite number of levels of para-
sites that we observe in this model appears to be related
to a transition that occurs in models for the growth of in-
terfaces by deposition of particles on surfaces [14 -- 16]. In
these models, the interface is modelled as a lattice, with
Figure 6: (a) Density of sites with n ≥ m, as a function of
level m, in the stacked CP for decreasing death rates dn =
f nd0, with d0 = 1, bn = 5 (solid lines) and for increasing
birth rates bn = f −nb0, with b0 = 5, dn = 1 (dashed
lines). In the former case, the exact CP predicts a nonzero
density as m → ∞ only for f ' 0.67.
(b) Density of
the mth level in the limit m → ∞ with bn = 5, dn =
f nd0, f = 0.75 as a function of host death rate d0 from
simulations (circles) and as predicted by a standard CP
(triangles) with λm→∞
= b(1 − f )/d0. The simulation
data is obtained by measuring the plateau values in plots
like panel (a), for m values up to 20. The mapping to a
standard CP predicts that only a finite number of levels
are sustainable for d0 > d∗
0 ≈ 0.758.
eff
a given height at each lattice site. According to the bal-
listic deposition rule for surface growth [18], particles fall
onto the lattice from above and only stick if the neigh-
bouring lattice site already contains a particle. In some
models, particles can also desorb from the surface [16]. In
the stacked CP model, we can think of the label m of a
given lattice site as corresponding to the local height of
the interface. The transmission of higher-level parasites
to a site with label m, from a neighbour with n > m then
corresponds to ballistic deposition, while death of a host
and loss of parasites loosely correspond to desorption of
particles. This apparent mapping is intriguing because
these models for surface growth show a roughening tran-
sition [14 -- 16]: when the deposition rate is low, the in-
terface remains smooth, that is, the width of the surface
layer remains finite in the thermodynamic limit, whereas
when the deposition rate is high, the surface layer grows
and roughens over time, that is, arbitrarily large differ-
ences in surface height can arise in the thermodynamic
limit. These two cases correspond to finite and infinite hi-
erarchies of parasites in our stacked CP model. Although
the mapping to the surface growth models is not exact,
one expects to see the same phenomenology, e.g. in terms
of critical exponents, for the transitions. For example a
generic model for coupled directed percolation processes
with unidirectional coupling between adjacent levels [25]
can show a different β exponent at different levels when
the critical points coincide. It would be interesting to see
if this is true for models such as the one described here
which show bidirectional coupling between all levels.
p-5
S. J. Court et al.
Discussion. -- In this paper, we have presented a sim-
ple model for the dynamics of long-lived infections in spa-
tially structured populations. The model is a stacked con-
tact process (CP), in which the host population undergoes
a standard CP, and a secondary CP representing the par-
asite takes place on the dynamically changing network of
lattice sites occupied by the host. We find that the cou-
pling between the dynamics of the host population and of
the infection leads to non-trivial effects, including a non-
monotonic dependence of the parasite prevalence on the
host population's turnover rate, for fixed host population
size. This exposes a connection between spatial and tem-
poral scales, since host population turnover affects spatial
clustering of infected organisms which in turn affects the
prevalence of the infection within the population. We have
found that an improved mean-field theory that takes local
correlations into account (the ordinary pair approximation
[19]) fails to reproduce this nonmonotonic behaviour. This
suggests that a deep understanding of the fluctuations in
the host CP is required to predict the steady-state prop-
erties of the parasite population. It will be interesting in
future to investigate how the nature of the CP phase tran-
sition for the infection dynamics is affected by the dynam-
ics of the underlying host population. Since disorder can
alter critical exponents and lead to new dynamics such as
Griffiths phases and activated scaling [20, 21], we envisage
that the secondary CP may show fundamentally different
behaviour to the standard CP.
Inspired by natural examples of hyperparasitism, or
"parasites on parasites", we have also investigated the
properties of a multi-level contact process. For this model,
we show that the average density of the m-th (and above)
level of the stacked CP maps onto that of a standard,
one-level CP, if the birth rates at all levels are equal. We
find that a phase transition separates two very different
behaviours of the model: maintenance of an infinite hier-
archy of levels of parasites, versus collapse to zero density
at a finite number of levels. In light of the link to models
of surface growth it will be interesting to determine more
fully the properties of the parasite maintenance transition
that we observe in the multi-layer stacked CP.
Finally, we note that the stacked CP model also presents
the possibility for a novel study of disorder in contact
processes. Classic models for disordered systems usually
create disorder by removing a fraction of sites from the
dynamics [22] or by allowing different creation or annihi-
lation rates for different sites [23]. This disorder is often
quenched - i.e. time-invariant. In our model, disorder in
the secondary CP (and higher CPs if present) arises nat-
urally from the underlying dynamics of the primary CP,
and can be controlled by varying the rates. This natu-
rally arising disorder is not fixed but varies in time and
space (in a fundamentally different process than the mo-
bile disorder of Ref. [24]). Future studies of the effects of
this natural disorder on the properties of the secondary CP
may uncover new principles of CP dynamics on disordered
lattices.
Acknowledgements. -- We thank Martin Evans for
useful discussions. SJC was funded by the Carnegie Trust
for the Universities of Scotland, RAB by an RCUK Aca-
demic Fellowship and RJA by a Royal Society University
Research Fellowship.
References
[1] Harris T., Ann. Probab. , 2 (1974) 969.
[2] Marro J. and Dickman R., Nonequilibrium Phase Tran-
sitions in Lattice Models (Cambridge University Press,
Cambridge) 1999.
[3] Henkel M., Hinrichsen H. and Lubeck S., Non-
equilibrium phase transitions, Volume 1: Absorbing phase
transitions (Springer, Dordrecht) 2008.
[4] Liggett T., Interacting Particle Systems (Springer, New
York) 1985.
[5] Summers D. K., The Biology of Plasmids (Blackwell Sci-
ence, UK) 1996.
[6] Smith J. and Dunn A., Parasitol. Today , 7 (1991) 14648.
[7] Lipsitch M., Nowak M., Ebert D. and May R., Proc.
R. Soc. B , 260 (1995) 321.
[8] Jones E., White A. and Boots M., Proc. R. Soc. B ,
278 (2011) 863.
[9] Busenberg S., Cooke K. L. and Pozio M. A., J. Math.
Biol. , 17 (1983) 305.
[10] Schinazi R., Math. Biosci. , 168 (2000) 1.
[11] Sullivan D. and Volkl W., Annu. Rev. Entomol. , 44
(1999) 291.
[12] Milgroom M. and Cortesi P., Ann. Rev. Phytopathol.
, 42 (2004) 311.
[13] Swinton J. and Gilligam C., Proc. Roy. Soc. Lond. B
, 266 (1999) 437.
[14] Alon U., Evans M. R., Hinrichsen H. and Mukamel
D., Phys. Rev. Lett. , 76 (1996) 2746.
[15] Alon U., Evans M., Hinrichsen H. and Mukamel D.,
Phys. Rev. E , 57 (1998) 4997.
[16] Blythe R. A. and Evans M. R., Phys. Rev. E , 64 (2001)
051101.
[17] Bortz A., Kalos M. and Lebowitz J., J. Comput.
Phys. , 17 (1975) 10.
[18] Meakin P., Ramanlal P., Sander L. M. and Ball
R. C., Phys. Rev. A , 34 (1986) 5091.
[19] Sato K., Matsuda H. and Sasaki A., J. Math. Biol. ,
32 (1994) 251.
[20] Griffiths R. B., Phys. Rev. Lett. , 23 (1969) 17.
[21] Vojta T. and Dickison M., Phys. Rev. E , 72 (2005) .
[22] Moreira A. G. and Dickman R., Phys. Rev. E , 54
(1996) R3090.
[23] Noest A. J., Phys. Rev. Lett. , 57 (1986) 90.
[24] Dickman R., J. Stat. Mech.: Theor. Exp. ,
(2009)
P08016.
[25] Tauber U. C., Howard M. J. and Hinrichsen H.,
Phys. Rev. Lett. , 80 (1998) 2156.
p-6
|
1603.07182 | 1 | 1603 | 2016-03-19T07:46:58 | Dark states in quantum photosynthesis | [
"physics.bio-ph",
"quant-ph"
] | We discuss a model of quantum photosynthesis with degeneracy in the light-harvesting system. We consider interaction of excitons in chromophores with light and phonons (vibrations of environment). These interactions have dipole form but are different (are related to non-parallel vectors of "bright" states). We show that this leads to excitation of non-decaying "dark" states. We discuss relation of this model to the known from spectroscopical experiments phenomenon of existence of photonic echo in quantum photosynthesis. | physics.bio-ph | physics |
Dark states in quantum photosynthesis
S.V.Kozyrev, I.V.Volovich
Steklov Mathematical Institute
August 9, 2018
Abstract
We discuss a model of quantum photosynthesis with degeneracy in the light-harvesting
system. We consider interaction of excitons in chromophores with light and phonons (vi-
brations of environment). These interactions have dipole form but are different (are related
to non -- parallel vectors of "bright" states). We show that this leads to excitation of non-
decaying "dark" states. We discuss relation of this model to the known from spectroscopical
experiments phenomenon of existence of photonic echo in quantum photosynthesis.
1
Introduction
Open quantum systems were investigated in many works, in particular, the stochastic limit method
was developed [1]. Degenerate open systems (when the behavior of quantum systems is more
complex) were discussed in particular in [2].
In this paper we consider a model of quantum photosynthesis (excitation, transport and ab-
sorption of excitons in light-harvesting systems) as a three level quantum system with energy levels
ε0, ε1, ε2, ε0 < ε1 < ε2, which interacts with three quantum fields (the reservoirs). The upper level
ε2 (which corresponds to one exciton states in the system of chromophores) is degenerate. The
two lower levels 0i, 1i are non degenerate and describe the state "exciton in the reaction center"
(state 1i with energy ε1) and the state without excitons 0i with energy ε0. Three possible tran-
sitions between the energy levels are paired to different reservoirs, which describe correspondingly
light, phonons (vibrations of protein matrix) and absorption of excitons in the reaction center.
Here the light can be coherent.
We consider the dynamics of open quantum systems with dissipation [3], i.e.
interaction
with any of the reservoirs is described by a generator for density matrices in the Lindblad form,
and for the light reservoir we also have an additional term in the generator which describes the
interaction with laser. Interaction of degenerate systems with fields generate the so called dark
states (or dark -- state polaritons) known in quantum optics [4], [5]. Moreover for different reservoirs
the corresponding spaces of bright and dark states can be different.
We conjecture that the interactions of chromophores with photons and phonons are different
and the bright photonic states generated by interaction with light will contain a dark component
for interaction with phonons. States of this kind will not decay in the process of transport of
excitons. We discuss the relation of these non-decaying dark states and known from experiments
with photonic echo [6], [7] quantum coherencies in photosynthesis systems.
1
Dynamics in the model under consideration is described by a sum of three generators --
photonic, phononic and the generator of absorption of excitons. We consider the following scheme
of computation of manipulation of quantum states of the system.
1) First, we consider the approximation of strong light and weak transport. In this approxi-
mation we take into account only the photonic generator and neglect the effects of transport and
absorption.
2) Second, we switch off the light and for the obtained at the previous stage photonic stationary
state we consider joint action of generators of transport and absorption. After this the upper level
of the system will contain only the part of the state which is dark with respect to the transport
generator.
3) For the obtained at stage 2 state we consider interaction with a coherent light, which allows
to investigate the possibility of photonic echo for this state.
In non degenerate case (when energy level ε2 is non degenerate) the considered in this paper
non-equilibrium quantum system was discussed in [8]. The flow of excitons in the system was
computed and excitation of quantum coherencies for interaction with laser was considered. When
the laser is switched off these coherencies (in non degenerate case) decay due to interaction with
phonons. Quantum dynamics in the stochastic limit in presence of laser field was studied in [9].
Properties of antibunched light are discussed for instance in [10].
In paper [11] effects of degeneracy in exciton transport were investigated by the stochastic
limit method. It was shown that, using constructive interference, it is possible to achieve quantum
amplification of exciton transport (the supertransport effect). The possibility of excitation of non-
decaying dark states in a degenerate system can be considered as a side effect of the supertransport.
In [12] dark states in photosynthetic systems were studied experimentally.
In [13] application of quantum methods to computations and biology was discussed. In [14]
dark states in photosynthesis were considered but dark states in this paper were considered in a
different way compared to the present paper (as collective states of excitons, photons and phonons),
in particular interactions between excitons and photons, excitons an phonons in [14] are similar
(in this case the effect considered in the present paper does not take place).
The exposition of the present paper is as follows.
In section 2 the Hamiltonian of light-
harvesting system interacting with three reservoirs (photons, phonons and absorption) is described.
In section 3 the corresponding generators of dynamics of density matrix of the system are con-
sidered (there are three such generators which describe excitation, transport and absorption of
excitons). In section 4 bright, dark and off-diagonal states for the mentioned generators are dis-
cussed. In section 5 preparation of quantum states by interaction with light (including laser field)
is considered. In section 6 the state obtained in section 5 is subjected to transport and absorption
(when the light is switched off). In section 7 we consider the interaction of the obtained at the
previous stage dark state of the system with laser field and discuss the relation to photonic echo
experiments.
2
2 Hamiltonian of the model
We consider a system with three energy levels ε0 < ε1 < ε2, where the upper level is degenerate,
with the Hamiltonian
N
HS = ε00ih0 + ε11ih1 + ε2
Xj=2
jihj.
(1)
This Hamiltonian describes a light-harvesting system, 0i corresponds to a state without ex-
citons, 1i is a state "exciton in the reaction center", ji correspond to one-exciton states of
chromophores.
System interacts with three quantum fields (reservoirs) in a dipole way. Transitions between
the levels with energies ε0 and ε2 (in particular creation of excitons) are related to interaction with
light (Bose field in the Gibbs state with the temperature β−1
em = 6000K, or laser field with the
frequency ε2 − ε0), transitions between the levels ε2 and ε1 (transport of excitons to the reaction
center) are related to interaction with phonons, or vibrations of protein matrix (described by
the Bose field with the temperature β−1
ph = 300K), and transitions between the levels ε1 and ε0
(absorption of excitons in the reaction center) are described by interaction with the sink reservoir
in the Fock state (i.e. reservoir with the zero temperature).
Thus we have three reservoirs described by Hamiltonians of quantum Bose fields Hem (light, or
electromagnetic field), Hph (phonons, or vibrations of protein matrix), Hsink (sink, or absorption
of excitons in the reaction center), each of the reservoir Hamiltonians has the form
HR = ZR3
ωR(k)a∗
R(k)aR(k)dk,
where R = em, ph, sink enumerate the reservoirs, ωR is the dispersion of the Bose field aR.
Each of the reservoirs is in a mean zero Gaussian state with the quadratic correlator
Here NR(k) (number of the field quanta with wave number k) is equal to (for the inverse
ha∗
R(k)aR(k′)i = NR(k)δ(k − k′).
temperature βR)
The full Hamiltonian has the form
NR(k) =
1
eβRωR(k) − 1
.
(2)
H = HS + Hem + Hph + Hsink + λ (HI,em + HI,ph + HI,sink) ,
where the interaction Hamiltonians HI,em, HI,ph, HI,sink have dipole forms and are given by for-
mulae (3), (4), (5) below.
Interaction of the system with light is described by the Hamiltonian
HI,em = Aemχih0 + A∗
em0ihχ,
A∗
em = ZR3
gem(k)a∗
em(k)dk,
(3)
where the bright photonic state χ belongs to the level with energy ε2, this state can be taken
normalized kχk = 1, gem(k) is the form -- factor of the field.
3
Transport of excitons to the reaction center is related to interaction of the system with phonons
HI,ph = Aphψih1 + A∗
ph1ihψ,
A∗
ph = ZR3
gph(k)a∗
ph(k)dk,
(4)
here the bright phononic state ψ with the energy ε2 is normalized kψk = 1.
Vectors ψ and χ belong to the same degenerate level with energy ε2 (corresponding to excitons
on chromophores). Crucial feature of the model under consideration is as follows -- vectors ψ and
χ are non-parallel.
Absorption of excitons in the reaction center is described by interaction with the additional
field of sink (with the zero temperature)
HI,sink = Asink1ih0 + A∗
sink0ih1,
A∗
sink = ZR3
gsink(k)a∗
sink(k)dk.
(5)
Remark. Usually in papers on photosynthesis some different Hamiltonian is used. In particular
in this Hamiltonian the part (1) is non-diagonal and contains terms corresponding to the dipole
interaction of excitons, the interaction Hamiltonian is also different. Here we consider the result of
diagonalization of the system Hamiltonian HS (transition to the so called "global" basis), i.e. in
our notations the states jiof the system correspond to the states from the "global" basis. Relation
between the "global" and "local" approaches in theory of open quantum systems was discussed in
[15].
3 Generators of the dynamics
For investigation of the model we will use the method of quantum stochastic limit [1], [2]. In
this limit dynamics of reduced density matrix of a system interacting with reservoir is generated
by operator in the Lindblad form, see formulae (7), (12), (13) below. Quantum dynamics in the
stochastic limit in presence of a laser field was discussed in [9].
For the considered model with three reservoirs the dynamics will be generated by a sum of
three generators
d
dt
ρ(t) = (θem + i[·, Heff] + θph + θsink) (ρ(t)).
We will manipulate quantum states of the system by switching different generators in this
formula. Actually it is possible to manipulate only with the photonic generator θem + i[·, Heff].
of the dissipative Lindblad term and the term related to coherent field [1, 2]:
Creation and annihilation of excitons are described by the photonic generator equal to a sum
θem(ρ) = 2γ−
Lem = θem + i[·, Heff], Heff = s(χih0 + 0ihχ),
2{ρ,χihχ}(cid:19) − iγ−
2{ρ,0ih0}(cid:19) + iγ+
re,em(cid:18)hχρχi0ih0 −
re,em(cid:18)h0ρ0iχihχ −
+2γ+
im,em[ρ,0ih0].
1
1
4
s ∈ R.
im,em[ρ,χihχ]+
(6)
(7)
The constants γ have the form
γ+
re,em = πZ gem(k)2δ(ωem(k) − ε2 + ε0)Nem(k)dk,
γ+
γ−
re,em = πZ gem(k)2δ(ωem(k) − ε2 + ε0)(Nem(k) + 1)dk,
im,em = −Z gem(k)2 P.P.
im,em = −Z gem(k)2 P.P.
ωem(k) − ε2 + ε0
Nem(k)dk,
1
1
(Nem(k) + 1)dk.
γ−
ωem(k) − ε2 + ε0
Here the function Nem(k) is given by (2), β−1
em = 6000K for the sun light.
For the laser field Nem(k) = 0, this implies γ+
re,em = γ+
im,em = 0, but γ−
re,em, γ−
dissipative part of the generator is non zero even for coherent field).
Transport of excitons is described by the phononic generator, here β−1
ph = 300K
im,em 6= 0 (i.e.
(8)
(9)
(10)
(11)
(12)
(13)
θph(ρ) = 2γ−
1
re,ph(cid:18)hψρψi1ih1 −
re,ph(cid:18)h1ρ1iψihψ −
2{ρ,ψihψ}(cid:19) − iγ−
2{ρ,1ih1}(cid:19) + iγ+
1
+2γ+
Absorption of excitons is described by the sink generator with β−1
im,ph[ρ,ψihψ]+
im,ph[ρ,1ih1].
sink = 0K
θsink(ρ) = 2γ−
re,sink(cid:18)h1ρ1i0ih0 −
1
2{ρ,1ih1}(cid:19) − iγ−
im,sink[ρ,1ih1].
Here coefficients γ+ are equal to zero since the temperature is zero.
Constants γ in (12), (13) are given by formulae analogous to (8), (9), (10), (11) with differ-
ent Bohr frequencies (differences of energies of the levels), dispersions and temperatures of the
reservoirs (i.e. we substitute these parameters by the corresponding for phonons and sink).
4 Bright, dark and off-diagonal states
For Lindblad generators of the form (7), (12), (13) one can consider expansion of the space of
states of the system in a sum of orthogonal subspaces of bright, dark and off-diagonal states.
Bright and dark states were extensively discussed in quantum optics, see [4], [5]. We use here the
approach and notations of [11].
These subspaces depend on the generator and are different for different generators. Let us
discuss the photonic generator (7). Generator θem describes creation and annihilation of excitons
by interaction of chromophores and electromagnetic field. Bright mixed states for this generators
are linear combinations of matrices
Below we usually will ignore the positivity and normalization for states, i.e. we will talk about
matrices instead of states (even if we will use the term state).
0ih0,
χihχ.
5
Pure bright state of the upper level (with energy ε2) is the vector χi in (3). Pure dark states
Mixed dark states are matrices B which give zero when multiplied by any mixed bright state
of the upper level are vectors φi orthogonal to the pure bright state: φ⊥χ.
A, i.e.
AB = BA = 0.
In the case of the generator θem these states are linear combinations of matrices
φihφ′,
1ih1,
φih1,
1ihφ, φ⊥χ, φ′⊥χ.
Off-diagonal matrices are matrices C orthogonal to all bright matrices A and all dark matrices
B with respect to the scalar product (·,·) = tr(··), i.e.
tr(CA) = tr(CB) = 0.
This subspace contains matrices corresponding to transitions between bright and dark sub-
spaces, and between levels in the bright subspace, for the generator θem the off-diagonal subspace
is generated by matrices
χih0,
χihφ,
χih1,
1ih0,
φih0, φ⊥χ
and conjugated.
Dark states described above are stationary with respect to the dynamics generated by θem. The
bright space corresponds to processes of creation and annihilation of excitons (for the generator
θem) and to process of transport of excitons (for the generator θph). Off-diagonal matrices decay
which corresponds to the decoherence phenomenon.
In presence of a coherent field (generator
i[·, Heff]) off-diagonal matrices can be created.
Described expansion of the state of matrices depends on the generator and for the phononic
generator θph will be completely different. In particular matrix 1ih1 is dark for θem and bright for
θph, and matrix χihχ is bright for θem and contains a combination of bright, dark and off-diagonal
terms for θph (since the phononic pure bright vector ψ is not parallel to the photonic pure bright
vector χ).
Let us discuss also bright, dark and off-diagonal matrices for the phononic generator θph given
by (12). The pure bright vector of the upper level is given by ψi in (4). Mixed bright matrices
for θph are linear combinations of
The space of mixed dark states is generated by matrices
1ih1,
ψihψ.
ηihη′,
0ih0,
ηih0,
0ihη,
η⊥ψ, η′⊥ψ.
The off-diagonal space is a linear span of matrices
ψih1,
ψihη,
ψih0,
0ih1,
ηih1,
η⊥ψ
and conjugated.
6
5 Stage 1: light is switched on
Let us discuss the process of manipulation of quantum states of light-harvesting system. The
initial state is the density matrix without excitons
ρ0 = 0ih0.
We apply to this initial state the dynamics generated by the photonic generator Lem = θem +
i[·, Heff] of the form (6) where the light contains also coherent component. This dynamics puts
the system in a stationary state of the form
ρ1 = ρ000ih0 + ρχχχihχ + ρχ0χih0 + ρ0χ0ihχ,
(14)
where
ρ00 =
re,em − s2
γ−
2 (cid:16) 1
µ20
+ 1
µ02(cid:17)
γ+
re,em + γ−
µ20
re,em − s2(cid:16) 1
re,em − γ+
γ−
re,em − s2(cid:16) 1
re,em
ρχ0 =
is
µ20
γ+
re,em + γ−
,
ρχχ =
γ+
re,em + γ−
+ 1
µ02(cid:17)
,
µ20
µ02(cid:17)
+ 1
µ20 = −γ−
re,em − γ+
re,em − γ+
µ02 = −γ−
ρ0χ = −
is
µ02
γ+
re,em + γ−
re,em + iγ−
re,em − iγ−
im,em + iγ+
im,em − iγ+
im,em,
im,em.
+ 1
µ02(cid:17)
µ20
re,em − s2(cid:16) 1
re,em − γ+
γ−
re,em − s2(cid:16) 1
re,em
µ20
re,em − s2
γ+
2 (cid:16) 1
µ20
+ 1
µ02(cid:17)
,
(15)
+ 1
µ02(cid:17)
,
(16)
(17)
(18)
One can say that we used the approximation of strong light and ignore the contributions to
the dynamics from the phononic generator θph and the sink generator θsink.
In particular, if there is no coherent field then off-diagonal elements of the matrix ρ1 vanish
ρχ0 = ρ0χ = 0 and diagonal elements ρ00, ρχχ are given by the Gibbs state satisfying
ρχχ
ρ00
= e−βem(ε2−ε0).
In the limit of strong coherent field s → ∞ we get for (15), (16)
ρ00 = ρχχ =
1
2
,
ρχ0 = ρ0χ = 0.
For the photonic generator Lem = θem + i[·, Heff] there exist also dark stationary states which
are linear combinations of matrices
φihφ′,
φih1,
1ihφ,
1ih1,
φ, φ′⊥χ.
Thus addition of a coherent field does not change the space of dark states.
Proof. Dynamics with the generator Lem and the initial sate ρ0 = 0ih0 belongs to the space of
matrices of a form (14). In the space of matrices (14) the matrix of the generator Lem is as follows
Lem
ρχχ
ρ00
ρχ0
ρ0χ
=
2γ+
−2γ−
re,em;
re,em;
is; −is
re,em; −2γ+
2γ−
re,em; −is;
is
0
µ20;
−is;
is;
is;
−is;
0;
µ02
ρχχ
ρ00
ρχ0
ρ0χ
.
7
For the stationary state the third equation of the system Lemρ = 0 implies
is(ρχχ − ρ00) = −µ20ρχ0,
i.e.
(ρχχ − ρ00).
Substitution of these ρχ0, ρ0χ in the first equation of Lemρ = 0 gives
(ρχχ − ρ00),
ρχ0 = −
ρ0χ =
is
µ20
is
µ02
(cid:18)2γ+
re,em − s2(cid:18) 1
µ20
+
1
µ02(cid:19)(cid:19) ρ00 = (cid:18)2γ−
re,em − s2(cid:18) 1
µ20
+
1
µ02(cid:19)(cid:19) ρχχ.
The normalization condition ρ00 + ρχχ = 1 implies (15), (16).
It might seem that the described stationary state does not depend on the degeneracy,
Remark.
in particular the off-diagonal element (16) is similar for both cases with and without degeneracy.
This formula follows from the normalization kχk = 1 in (3) which is always possible: if kχk 6= 1
then one can the corresponding normalization add to the expression for the field in (3):
HI,em = Aemχih0 + A∗
em0ihχ = kχkAem
1
kχkχih0 + kχkA∗
em
1
kχk0ihχ.
Therefore the normalization kχk = 1 means that the parameter s is multiplied by kχk and
values γ±
em are multiplied by kχk2. This transformation gives
γ−
re,em − γ+
re,em − s2
γ+
re,em + γ−
kχkµ20
ρχ0 =
is
re,em
kχk2 (cid:16) 1
µ20
.
+ 1
µ02(cid:17)
6 Stage 2: light is switched off
The second stage of manipulation of quantum states of light-harvesting system is as follows: we
switch off the light, i.e. we use the obtained at the previous step state ρ1 given by (14), (15), (16)
as the initial state for dynamics with the generator θph + θsink (sum of phononic generator and
sink generator).
Let us consider the expansion of the bright photonic state
in the sum of contributions parallel and orthogonal to the bright phononic state ψ
χ = χ0 + χ1,
χ0kψ, χ1⊥ψ
χ0i = hψ, χiψi = ψihψχi,
χ1i = (1 − ψihψ)χi.
Let us substitute this expansion in expression (14), (15), (16) for the stationary photonic state
and apply the generator θph + θsink. Discussion of section 4 of dynamics in spaces of bright, dark
and off-diagonal matrices implies that all terms in the expansion which contain χ0 will decay since
the transport generator θph will transfer excitons to the reaction center where excitons will be
absorbed. Hence the system will tend to the stationary state of the form
ρ2 = ρ000ih0 + ρχχχ1ihχ1 + ρχ0χ1ih0 + ρ0χ0ihχ1,
(19)
8
where ρχχ, ρχ0, ρ0χ are given by (15), (16), and ρ00 is given by the condition that trace of density
matrix is equal to one
ρ00 = 1 − kχ1k2ρχχ.
In the model under consideration the obtained state (19) can exist infinite time (in reality
the lifetime should be finite but long). If the bright states for photons and phonons coincide (i.e.
χ1i = 0) then we will get ρ2 = 0ih0. In general case of non-parallel ψ, χ the obtained state ρ2
will contain non-decaying dark part. This part can be observed in spectroscopical experiments,
see the next section.
Remark. Vectors ψ, χ are norm one vectors in multidimensional space. Random independent
normed vectors in a multidimensional space are close to orthogonal (since measure of a unit sphere
is concentrated near the equator of this sphere).
On the other hand, biological systems undergo selection pressure, for the present model this
is the selection pressure to increase effectiveness of exciton transport. This selection pressure will
aim to make vectors χ and ψ parallel (in this case the transport will be most effective). Therefore
the angle between χ and ψ will be related to competition between entropy and selection.
7 Stage 3: light is switched on
At the third stage of manipulation of quantum states of excitons we subject the obtained at the
previous step state ρ2 of the form (19) to interaction with coherent light. We ignore transport
and absorption (for example the light strong and we consider small times). We will show that in
this situation it is possible to observe photonic echo.
Spectroscopy is the application of dynamics with the generator i[·, Heff] to the off-diagonal part
of density matrix proportional to χih0− 0ihχ, see below. Thus we are interested in the term in
(19) of the form
(20)
ρχ0χ1ih0 + ρ0χ0ihχ1,
where the matrix elements ρχ0, ρ0χ are given by (16). In particular this term can be non-zero only
if s 6= 0, i.e. the state (14) should be excited by light with coherent component.
Let us recall that χ1 is the orthogonal complement to projection of χ (bright photonic vector)
to ψ (bright phononic vector). We define χ2 as projection of χ1 to χ (i.e. χ2 is parallel to χ), then
Applying i[·, Heff] to (20) we will get non-zero contribution only for the following matrix (since
χ2i = (cid:0)1 − hψ, χi2(cid:1)χi.
the contribution from orthogonal complement to χ2i vanish)
ρχ0χ2ih0 + ρ0χ0ihχ2 =
= Re ρχ0(cid:0)1 − hψ, χi2(cid:1) (χih0 + 0ihχ) + iIm ρχ0(cid:0)1 − hψ, χi2(cid:1) (χih0 − 0ihχ) ,
where ρχ0 is given by (16). Non-zero contribution to interaction with the laser here gives the
second term, equal to
ρ3 = isRe(cid:18) 1
µ20(cid:19)
γ+
re,em + γ−
re,em
γ−
re,em − γ+
re,em − 2s2Re(cid:16) 1
µ20(cid:17) (cid:0)1 − hψ, χi2(cid:1) (χih0 − 0ihχ) .
(21)
9
In the regime under consideration (s 6= 0) this term is non-zero. This term is largest when the
laser amplitude is chosen as follows
s2 = −
re,em
re,em + γ−
γ+
µ20(cid:17)
2Re(cid:16) 1
=
1
2 (cid:0)γ−2
re,em + γ+2
re,em + γ−2
im,em + γ+2
im,em(cid:1) ,
In this case
iIm ρχ0 = −
i
2√2
re,em − γ+
γ−
re,em
re,em + γ−2
re,em + γ+2
qγ−2
.
im,em + γ+2
im,em
If the state ρ1 was excited by purely coherent light when Nem(k) = 0, hence
γ+
re,em = γ+
im,em = 0,
γ−
re,em = πZ gem(k)2δ(ωem(k) − ε2 + ε0)dk,
im,em = −Z gem(k)2 P.P.
1
ωem(k) − ε2 + ε0
γ−
dk,
then Im ρχ0 is maximal. If we ignore the shift γ−
im,em we get the simple expression
ρχ0 = −
i
2√2
.
Summary. We have shown that the model of light-harvesting complex as a degenerate system
with absorption and interaction with photons and phonons describes excitation of dark states
which will have long lifetime and will be visible in spectroscopic experiments.
Earlier it was shown [11] that degeneracy can be used for quantum amplification of exciton
transport (the supertransport effect). Results of the present paper show that in this case, as a side
effect of the supertransport, we will obtain coherent dark states with long lifetime. This result
can be discussed in relation of phenomenon of quantum photosynthesis [6], [7] and experimental
observation of dark states in photosynthetic systems [12].
Remark. We have discussed manipulations with quantum states, in particular with dark states:
excitation of the states by coherent fields and using of Lindblad dissipation for projections of the
states to some subspaces of matrices; since the bright vectors for different fields (χ and ψ in the
present model) can be non-parallel the manipulations under consideration are nontrivial. The
considered manipulations using degeneracy and different ways of interaction with the degenerate
energy level of the system are of very special form.
Manipulations of quantum states can be used for quantum computations [13]. Long lifetime
of dark states could help to avoid decoherence in quantum computers. For another proposal to
reduce decoherence in quantum computers by manipulating with macroscopic parameters see [16].
Different ideas of how to use photosynthetic quantum effects for computations are discussed in
[6], [17], [18].
It would be interesting also to discuss a possible connection of manipulations with quantum
states considered in this our work with those studied in quantum control theory, see for example
[19].
10
Quantum beats. Let us recall the model of interaction of two level system with laser. Dynamics
of the density matrix of a system with levels 0i, χi (where kχk = 1) is described by equation
d
dt
ρ(t) = i[ρ(t), H],
H = sσx = s(cid:18) 0 1
1 0(cid:19) ,
s ∈ R.
The basis of sigma -- matrices is
σz = χihχ−0ih0.
σ0 = χihχ+0ih0,
Operators σ0, σx are invariant with respect to the dynamics. Action of the generator of the
dynamics in the subspace
σy = i (−χih0 + 0ihχ) ,
σx = χih0+0ihχ,
is given by
ρ = r(χihχ − 0ih0) + ui (−χih0 + 0ihχ)
i[·, H](cid:18) r
u(cid:19) = (cid:18) 0
−2s
2s
0 (cid:19)(cid:18) r
u(cid:19) .
The corresponding dynamics reduces to oscillations in this two-dimensional space with Rabi
frequency Ω = 2s.
Acknowledgments. This research was funded by a grant from the Russian Science Foundation
(Project No. 14-50-00005).
References
[1] L. Accardi, Lu Yun Gang, I. Volovich, Quantum theory and its stochastic limit, Springer-
Verlag, Berlin, 2002
[2] L.Accardi, S.V.Kozyrev. Lectures on Quantum Interacting Particle Systems, in: QP-PQ:
Quantum Probability and White Noise Analysis - Vol. 14 "Quantum Interacting Particle
Systems", World Scientific Publishing, 2002.
[3] A. S. Holevo, Quantum systems, channels, information, DeGruyter, 2012.
[4] M.O. Scully, M.S. Zubairy, Quantum Optics, Cambridge University Prress, Cambridge, UK,
1997.
[5] M. Fleischhauer, M.D. Lukin, Dark-State Polaritons in Electromagnetically Induced Trans-
parency, Phys. Rev. Lett. 2000, V.84, P.5094, arXiv:quant-ph/0001094
[6] G.S. Engel, T.R. Calhoun, E.L. Read, T.K.Ahn, T. Mancal, Y.C. Cheng, R.E. Blankenship,
G.R. Fleming, Evidence for wavelike energy transfer through quantum coherence in photo-
synthetic systems. Nature, 2007, V.446. P.782 -- 786.
[7] G.D. Scholes, G.R. Fleming, A. Olaya-Castro, R. van Grondelle, Lessons from nature about
solar light harvesting. Nature Chem., 2011, 3, 763774.
11
[8] S.V.Kozyrev, A.Mironov, A.Teretenkov, I.V.Volovich, Three-level system interacting with
three reservoirs and quantum photosynthesis, To appear.
[9] L. Accardi, S.V.Kozyrev, A.N.Pechen, Coherent Quantum Control of Λ -- Atoms through the
Stochastic Limit, P.1-17, in: Quantum Information and Computing, QP-PQ: Quantum Prob-
ability and White Noise Analysis - Vol. 19, eds. L.Accardi, M.Ohya, N.Watanabe, 2006, World
Scientific. arXiv:quant-ph/0403100
[10] Igor V. Volovich, Cauchy -- Schwarz inequality-based criteria for non-classicality of sub-Poisson
and antibunched light, Phys. Lett. A, 380:1-2 (2016), 5658.
[11] I. Ya. Aref'eva, I. V. Volovich, S. V. Kozyrev, Stochastic limit method and interference in
quantum many-particle systems, Theoretical and Mathematical Physics, 2015, 183:3, 782 -- 799
[12] M. Ferretti, R. Hendrikx, E. Romero, J. Southall, R.J. Cogdell, V.I. Novoderezhkin, G.D.
Scholes, R. van Grondelle, Dark States in the Light-Harvesting complex 2 Revealed by Two-
dimensional Electronic Spectroscopy, Scientific Reports, 6:20834, DOI: 10.1038/srep20834
[13] Masanori Ohya, Igor Volovich, Mathematical Foundations of Quantum Information and Com-
putation and Its Applications to Nano- and Bio-systems, Springer, 2011.
[14] Hui Dong, Da-Zhi Xu, Jin-Feng Huang, Chang-Pu Sun, Coherent excitation transfer via
the dark-state channel in a bionic system, Light: Science & Applications (2012) 1, e2;
doi:10.1038/lsa.2012.2
[15] A. S. Trushechkin, I. V. Volovich, Perturbative treatment of inter-site couplings in the local
description of open quantum networks, EPL, 113:3 (2016), 30005.
[16] I.V. Volovich, Models
of Quantum Computers
and Decoherence Problem,
arXiv:quant-ph/9902055.
[17] M. Mohseni, P. Rebentrost, S. Lloyd, A. Aspuru-Guzik, Environment -- assisted quantum walks
in photosynthetic energy transfer, J. Chem. Phys. 129, 174106 (2008).
[18] Gabor Vattay, Stuart A. Kauffman, Evolutionary Design in Biological Quantum Computing,
arXiv:1311.4688 [cond-mat.dis-nn].
[19] A.N. Pechen, N.B. Il'in, Existence of traps in the problem of maximizing quantum observable
averages for a qubit at short times, Proc. Steklov Inst. Math., 289 (2015), 213 -- 220.
12
|
1502.07082 | 2 | 1502 | 2015-10-22T03:48:04 | A New Model for the Collective Behavior of Animals | [
"physics.bio-ph"
] | We propose a new model in order to study behaviors of self-organized system such as a group of animals. We assume that the individuals have two degrees of freedom corresponding one to their internal state and the other to their external state. The external state is characterized by its moving orientation. The rule of the interaction between the individuals is determined by the internal state which can be either in the non-excited state or in the excited state. The system is put under a source of external perturbation called "noise". To study the behavior of the model with varying noise, we use the Monte-Carlo simulation technique. The result clearly shows two first-order transitions separating the system into three phases: with increasing noise, the system undergoes a phase transition from a dilute disordered phase to an ordered compact phase and then to the disordered dispersed phase. These phases correspond to behaviors of animals: uncollected state at low noise, flocking at medium noise and runaway at high noise, respectively. | physics.bio-ph | physics |
A New Model for Collective Behaviors of Animals
P. The Nguyena, V. Thanh Ngob∗ and H. T. Diepc
aDepartment of Natural Science, Duytan University,
K7/25 Quang Trung, Haichau, Danang, Vietnam
bInstitute of Physics - Vietnam Academy of Science and Technology,
10 Daotan, Ngoc Khanh, Badinh, Hanoi, Vietnam
cLaboratoire de Physique Th´eorique et Mod´elisation, Universit´e de Cergy-Pontoise, CNRS,
UMR 8089, 2 Avenue Adolphe Chauvin, F-95302 Cergy-Pontoise Cedex, France
We propose a new model in order to study behaviors of self-organized system such as a group
of animals. We assume that the individuals have two degrees of freedom corresponding one to
their internal state and the other to their external state. The external state is characterized by
its moving orientation. The rule of the interaction between the individuals is determined by the
internal state which can be either in the non-excited state or in the excited state. The system is put
under a source of external perturbation called "noise". To study the behavior of the model with
varying noise, we use the Monte-Carlo simulation technique. The result clearly shows two first-order
transitions separating the system into three phases: with increasing noise, the system undergoes
a phase transition from a dilute disordered phase to an ordered compact phase and then to the
disordered dispersed phase. These phases correspond to behaviors of animals: uncollected state at
low noise, flocking at medium noise and runaway at high noise, respectively.
I.
INTRODUCTION
The collective behavior of animals is a widely observed
phenomenon in various biological systems. It is one of
the main topics which have been extensively investigated
during the last two decades using methods borrowed from
many different areas of science including physics, ap-
plied mathematics and engineering. For recent papers,
the reader is referred to works of Couzin and cowork-
ers, of Albano and his group among others [2 -- 6].
In
these works, many aspects of animal groups have been
discussed, among which the collective motion and their
origin, the relation between individual and collective be-
haviors and the relation between group size and collective
decision-making.
The flocking is a behavior of some animal species where
they stay together in a group for social reasons. They de-
rive many benefits from this behavior including defence
against predators, easier collective moving, enhanced for-
aging success and higher success in finding a mate. When
they are faced with a danger such as predators, their nat-
ural instinct is to flee not to fight. They use their natural
herding instinct to bind together in a group for safety. All
individuals of the group will move away from the preda-
tor in the same direction and then stampede as fast as
they can when being under the predator's attack. If there
is no danger, then they are spread to find foods instead
of staying in the flocking state. Many experimental facts
and observations have been for example mentioned in the
review of Vicsek and Zafeiris [1]. Well-known examples
are found in populations such as large schools of fish [7]
or gatherings of birds [8]. Biologically, it is known that
the flocking behavior is advantageous for survival of a
∗Corresponding author, E-mail: [email protected]
population [9 -- 11]: reducing the risk of capture by preda-
tors, increasing higher mating efficiency, easier search for
food, efficient learning of external stimuli, and reducing
overall aggression [12 -- 15].
In 1987, Reynolds first suggested a simple model con-
sisting of three rules: separation, alignment, and cohe-
sion rules [16]. These rules describe the behavior of each
individual in interaction with other neighboring individ-
uals. All or some of the three rules were mathematically
expressed and then analyzed by Vicsek and his cowork-
ers [1, 17 -- 19]. They mainly focused on the transition
between coherently moving and runaway in a stampede.
The flocking behavior has been conventionally stud-
ied through simulation in two frameworks: population
(Eulerian or continuum models) and individuals (agents
or particle-based models) [14, 20, 21].
In the popula-
tion framework, the flock was collectively addressed while
flock-density was used as a key variable to present spatial
and temporal dynamics of aggregation frequently with
partial differential equations of advection-diffusion reac-
tion [22, 23]. In the individual framework, the flock of
agents has been simulated by using ordinary and stochas-
tic equations of motion to describe interactions among
agents [7, 17, 24 -- 26]. This approach attempted to repli-
cate naturally observed phenomena from not only animal
groups but also other self-propelled characteristics [27]
and to compare the evolved characteristics with those of
actual animal flocking [28] in order to better understand
possible mechanisms by which these characteristics may
have evolved.
The previous models have predicted with success flock-
ing behavior of animal groups at high noise. They have,
however, difficulties to explain uncollected states at very
low noise. This has motivated our present work: in addi-
tion to the interaction between neighboring animals, we
introduce an internal degree of freedom which indicates
whether an animal is in a non-excited or an excited state,
we will show that the collective behavior of animals in the
whole range of noise can be explained.
Section II is devoted to the description of the model.
Section III shows the Monte Carlo (MC) simulation re-
sults for testing our model. Conclusions are given in Sec.
IV.
II. THE MODEL
In biology, all the members of an animal group are
spread to find foods if there is no danger. In this situa-
tion, they are distributed in the space with a small con-
centration and out of alignment. So, we say the group
of animals is in a "uncollected" behavior. When the ani-
mals are faced with danger such as predators, they bind
together in a small area for safety with the same orienta-
tion and high concentration. This state is called "flock-
ing" state. Facing a danger, animals will move away from
the predator in the same direction and then stampede as
fast as they can when being under the predator's attack.
At the final stage, they are in a "runaway" state.
In order to study the phase transition behavior, we
have to map the group of animals into a physical system.
We consider an animal as a particle i with two degrees
of freedom: an external parameter σi characterizing the
animal orientation which depends on its interaction with
the others, and an internal parameter Si indicating either
it is in the non-excited (Si = 0) or in the excited (Si = 1)
individual state.
Let us define the internal degree of freedom. The inter-
nal state of a particle can be described by two levels: the
first level is characterized by a negative energy −ǫ which
describes the calm, unworried stable state, and the sec-
ond level by a positive energy +ǫ which expresses some
degree of anxiety. Introducing an external noise η which
plays the role of the temperature in statistical physics, we
can write the canonical probability of these two states as
[29]
p1 =
eǫ/η
z
, p2 =
e−ǫ/η
z
where z is the partition function given by
z = eǫ/η + e−ǫ/η = 2 cosh(ǫ/η)
(1)
(2)
Note that p1 > p2 because ǫ > 0. So, the "positive" ten-
dency towards tranquility is proportional to p1 − p2. The
total number of particles having this tranquility state
which is called non-excited state is thus given by
N0 = N (p1 − p2) = N (cid:20) eǫ/η − e−ǫ/η
z
(cid:21)
= N tanh(ǫ/η)
(3)
where N is the total number of particles in the system.
The number of excited particles is then
2
One sees in the above equations that in zero noise one has
N0 = N , namely there is as expected no excited particles.
While, when the noise goes to infinity, one has N0 = 0
and Ne = N : all particles are excited. At a given noise,
statistically one has a well-defined N0 and Ne with the
conservation N = N0 + Ne.
In the simulation which will be done below at a given
noise, we generate at random N0 according to Eq. (3).
The remaining particles in the system are thus in the
excited state.
Now, we put this system of individuals on the lattice
where each individual can move on 2D triangular lattice
of linear size L. The number of lattice sites should be
greater than that of individuals, i.e., L2 ≫ N . We denote
by σi the orientation of individual Si: σi is defined as in a
q-state Potts model, i.e. σi = 1, 2, . . . , q. For simplicity,
we consider q = 6, so the orientations σi = 1, 2, . . . , 6 of
individual Si can be defined by the vectors which connect
a site to its nearest neighbors (NN) with the following
angles measured from the x axis: ϕi = 0, π/3, . . . , 5π/3.
The interaction between nearest-neighboring animals is
given by the Hamiltonian
H = X<i,j>
Ki,j cos[π(σi − σj)/3],
(5)
where the sum P<i,j> is made over up to the third near-
est neighboring individuals Si and Sj. For simplicity,
Ki,j is assumed to take the form of the Lennard-Jones
potential:
Ki,j = 4J (cid:2)(r0/ri,j)12 − (r0/ri,j)6(cid:3) ,
where ri,j is the distance between two individuals. We
choose r0 = 0.89 in order that Ki,j ≃ J at ri,j = 1 which
is the lattice spacing, J being a positive constant. The
interaction between two individuals depends on their in-
ternal state: Ki,j = 0 if both individuals are non-excited
Si = Sj = 0, and Ki,j 6= 0 if otherwise. We empha-
size that the interaction rule we impose means that an
excited particle can interact with non-excited particles,
dragging them into a collective motion.
Let us note that, without the internal degree of free-
dom and without particle motions, the above Hamilto-
nian is the localized Potts clock model which has been
solved for q = 2, 3 and 4 [30]. The case of very large q has
been solved by Frohlich and Spencer [31]. However, for
q = 5, 6, ... there are not (yet) exact results. Of course,
our model is more complicated because the particles are
mobile with an internal degree of freedom.
III. THE PHASE-TRANSITION BEHAVIOR
We use the MC simulation technique to study the
above model. The main physical quantities such as the
order parameter Q and the concentration ρ are defined
in what follows.
Ne = N − N0
(4)
For Q, we have
where
with
Q =
qMmax − 1
q − 1
Mmax =
max(M1, M2, ..., Mq)
N
Mi = Xj
δσj ,i (i = 1, 2, ..., q)
(6)
(7)
(8)
where the sum on j is performed over all sites of the
system. One sees that in the ordered state where there
is only one kind of σj one has Q = 1. In the disordered
phase where all values of σj are equally present, namely
Mi = N/q for any i = 1, ..., q, one has Q = 0.
For the concentration ρ we have
ρ =
1
N
N
Xi=1
ni,
(9)
where ni is the number of NN individuals around Si. The
quantity ρ in Eq. (9) characterizes the spatial distribu-
tion of the individuals. The behavior of animals can be
adequately described by the two parameters Q and ρ.
In the simulations, we use ǫ = 0.04, J = 1.0 (taken as
the unit of energy) and N = 100, 400 and 900 with the
√N . For the initial configuration of
lattice size L = 4 ×
the system, we randomly generate the particle positions
on the lattice and take orientations of all individuals from
an uniform distribution. At each MC step, we randomly
choose N0 individuals (according to Eq. (3)), and set
them to be non-excited Si = 0. The remaining particles
are in the excited state Si = 1. Their position and ori-
entation are updated by Metropolis algorithm. At each
η, the equilibration time lies around 4 × 106 MC steps
per individual and we compute statistical averages over
8 × 106 MC steps per individual. Each time a particle
gets out of the lattice at one boundary, we put it into
the system by the other end using the periodic bound-
ary conditions in order to conserve the total number of
particles.
We plot in figures 1 and 2 the order parameter and
the concentration as a function of external noise for sev-
eral system sizes. It clearly shows the existence of three
phases which are separated by the two transitions at very
low noise and high noise.
In phase I at low noise, the
system is in the disordered phase with low concentra-
tion (Q ≈ 0 and ρ ≈ 0). This phase is equivalent to
the uncollected behavior of an animal group which is dis-
persed over the whole space because almost of them are
non-excited, it is called "free" phase with a few "con-
tacts" between the individuals. Phase III, at high noise,
is spatially sparse as the phase I, but the particles are
very mobile in all directions.
It describes the runaway
3
N = 100
N = 400
N = 900
(a)
(b)
0.1 0.2 0.3 0.4 0.5 0.6 0.7
Noise
r
e
t
e
m
a
r
a
p
r
e
d
r
O
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0
FIG. 1: Order parameter versus noise with the system sizes
N = 100 (circles), 400 (squares) and 900 (diamonds). The
insets show the enlarged scale at low (a) and high (b) noise.
6
5
4
3
n
o
i
t
a
r
t
n
e
c
n
o
C
2
1
0
0
0.1
0.2
0.3
0.4
Noise
N = 100
N = 400
N = 900
0.5
0.6
0.7
FIG. 2: Concentration versus noise with N = 100 (circles),
400 (squares) and 900 (diamonds).
behavior. At medium noise, the animals are in phase II
where they are compactly moving in an ordered phase
with Q ≈ 1 and the maximum concentration ρ ≈ 6. This
phase corresponds to the flocking behavior.
The spatial distribution of the individuals is also pre-
sented in Fig. 3 with snapshots taken at several values of
noise η = 0.0522, 0.1367, 0.5000 and 0.5667, at the final
stage of a MC simulation. The vectors indicate their po-
sition and moving orientation. One sees that Figs 3(a,d)
show that the system has the same distribution where
they are randomly distributed in the plane with differ-
ent orientations. However, the instantaneous snapshots
do not permit to see the difference between two phases:
the particles in phase I are almost immobile while they
are moving very fast in a disordered manner in phase III
(this difference is seen in videos). Fig. 3(b) shows the
flocking behavior, and Fig. 3(c) shows the distribution
of the system at the noise closes to the II-III transition
point.
At this stage, let us examine again Fig. 1:
it shows
discontinuities of the order parameter at both transitions
I-II and II-III, indicating a signature of a first-order tran-
sition. This abrupt behavior corresponds to the fact that
the animals immediately flock when they are faced with
danger, and they runaway when being under a predator's
attack. These behaviors are more likely a first-order tran-
(a)
(c)
η = 0.0522
η = 0.1367
(b)
(d)
r
e
t
e
m
a
r
a
p
r
e
d
r
O
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
4
η = 1.1126
η = 0.5844
0
0.2 0.4 0.6 0.8
Noise
1
1.2 1.4 1.6
η = 0.5000
η = 0.5667
FIG. 3: Snapshots at different values of noise with N = 100.
The arrow indicates the orientation of individual. See text for
comments.
(a)
η = 0.0768
η = 0.0736
η = 0.0723
N = 100
N = 400
N = 900
1
0.8
0.6
0.4
0.2
m
a
r
g
o
t
s
i
H
0
−2
−1.6
−1.2
−0.8
−0.4
0
(b)
η = 0.5414
η = 0.5745
η = 0.5844
N = 100
N = 400
N = 900
1
0.8
0.6
0.4
0.2
m
a
r
g
o
t
s
i
H
0
−2.5
−2
−1.5
Energy
−1
−0.5
0
FIG. 4: Histogram versus energy at low (a) and high (b)
noise at the transition points, with the system sizes N = 100
(solid), 400 (dotted) and 900 (dashed).
sition. To confirm the first-order transition we have per-
formed the energy histogram at the transition point. We
show in Fig. 4 the energy histogram at two transitions
I-II and II-III, for three system sizes. The double-peak
histograms are clearly shown for both transitions at low
noise (a) and high noise (b). The dip between the two
maxima becomes deeper with increasing size. Note that
FIG. 5: Order parameter versus noise with N = 100 for the
localized lattice model (diamonds) and lattice gas model (cir-
cles).
a "true" discontinuity happens only when the dip comes
down to zero. The distance between the two peaks is
then the latent heat. To see this, we need sizes much
larger than N = 900. But this is out of the scope of our
present purpose.
Let us discuss on the phase transition at high noise.
It is known that the q-state Potts model localized on
a lattice has a first-order transition with q ≥ 4 in two
dimensions. The clock Potts model localized on a lat-
tice was solved only for q ≤ 4 [30], though it was solved
for very large q [31]. The model treated in the present
paper corresponds to a mobile Potts model where par-
ticles can go from one lattice site to another. This was
never solved before. Note however that the transition II-
III corresponds not only to an orientational disordering
of the Potts parameter but also to the breaking of the
compactness of the flocking state. This is similar to a
melting transition where the solid phase melts into the
liquid phase. The melting has often a first-order char-
acter in three dimensions [32]. In two dimensions, it is
known that long-range solid ordering does not survive
at finite temperatures if the system has isotropic short-
range interactions according to Nelson and Halperin [33]:
at a first critical temperature, bound pairs of dislocations
formed at low temperatures are unbound, giving rise to
a phase with no translational ordering but with orien-
tational hexatic structure. The latter phase undergoes a
"critical" phase transition to the disordered phase. It was
however found by MC simulation that the second transi-
tion is not critical but it is a first-order two-dimensional
disclination melting [34]. Our results shown in Figs. 3
(c,d) and 4 indicate that our model undergoes a first-
order transition from the hexatic ordered phase to the
liquid phase, in agreement with earlier MC simulations
[34].
Finally, to compare our mobile Potts model with the
corresponding localized Potts model, we have simulated
the same system defined by (5) including the internal
degree of freedom on a lattice of size L2 = N , i.e. the
individuals are not allowed to move on the lattice. We
show in Fig. 5 results of both mobile and localized Potts
models: the transition temperature is about 0.5844 for
5
the mobile case, while it is η ≈ 1.1126 for the localized
Potts model.
IV. CONCLUSIONS
We proposed in this paper a simple model for studying
the behavior of animal groups as a function of an exter-
nal perturbation, called "noise", such as dangers coming
from predators. We showed that with increasing noise,
the system has three phases I, II and III separated by
two transitions, the first transition occurs at a low noise
and and the second one at high noise. The three phases
are disordered, ordered and melted phases which corre-
spond respectively to the following behaviors of animals:
uncollected state, flocking state and runaway state. Both
transitions from one phase to another are found to be of
the first-order.
This model is similar to the lattice gas model devel-
oped by Csah´ok et. al. [18] but in our model the align-
ment rules of the individuals depend on the internal state,
excited or non-excited, of its neighbors including itself.
Therefore, our model can be applied not only for study-
ing the phase transition behavior of animals at high noise
as in previous models, but also for analyzing the animal
behavior at low noise where animals are in a uncollected,
disordered phase.
This work was supported by the Nafosted (Vietnam
National Foundation for Science and Technology Devel-
opment), Grant No. 103.02-2011.55.
[1] T. Vicsek and A. Zafeiris, Phys. Reports 517, 71 (2012).
[2] I. D. Couzin, Trends in Cognitive Sciences 13, 36 (2009);
I. D. Couzin, J. Krause, N. R. Franks and S. A. Levin,
Nature 433, 513 (2005).
[3] G. Baglietto and E. V. Albano, Computer Physics Com-
munications 180, 527 (2009).
O. Shochet, Phys. Rev. Lett. 75, 1226 (1995).
[18] Z. Csah´ok and T. Vicsek, Phys. Rev. E 52, 5297 (1995).
[19] A. Czirok, H. E. Stanley, and T. Vicsek, Phys. Phys. A,
30, 137 (1996); A. Czir´ok, M. Vicsek, T. Vicsek, Physica
A 264, 299 (1999); A. Czir´ok and T. Vicsek, Physica A
281, 17 (2000).
[4] M. Aldana, H. Larralde and B. V´asquez, Internat. J. of
[20] D. Grunbaum and A. Okubo, Frontiers in Mathematical
Modern Physics B 23, 3459 (2009).
Biology, (S.A. Levin (Ed.), Springer, Berlin, 1994).
[5] G. P. Saracco, G. Gonnella, D. Marenduzzo and E. Or-
[21] Y. Inada and K. Kawachi, J. Theor. Biol. 214, 371
landini, Cent. Eur. J. Phys. 10, 1109 (2012).
(2002).
[6] H. Chat´e, F. Ginelli, G. Gr´egoire, F. Peruani and F. Ray-
[22] A. Mogilner and L. Edelstein-Keshet, Physica D 89, 346
naud, Eur. Phys. J. B 64, 451 (2008).
(1996).
[7] A. Huth and C. Wissel, Ecol. Modell., 75, 135 (1994).
[8] M. Maldonado-Coelho and M. A. Marini, Biol. Conserv.
[23] J. D. Murray, Mathematical Biology, (vols. 1 and 2,
Springer, New York, 2003).
116, 19 (2004).
[9] G. M. Werner and M. G. Dyer, Evolution of herding be-
havior in artificial animals, Proceedings of the Second
International Conference on Simulation of Adaptive Be-
havior, MIT Press, Cambridge, MA (1992).
[10] T. J. Pitcher (Ed.) and J. K. Parrish, Pitcher, The Be-
havior of Teleost Fishes, (Chapman & Hall, New York,
1993).
[11] W. Cresswell, Tringa tetanus. Anim. Behav. 47, 433
(1994).
[12] D. H. Cashing and F. R. Harden-Jones, Nature 218, 918
(1968).
[24] H. S. Niwa, J. Theor. Biol. 181, 47 (1996).
[25] Y. Tu, Phys. A 281, 30 (2000).
[26] J. Toner and Y. Tu, Phys. Rev. E 58, 4828 (1998).
[27] L. Spector, J. Klein, C. Perry, and M. Feinstein, GPEM
6, 111 (2005).
[28] A. J. Wood and G. J. Ackland, P. Roy. Soc. B-Biol. Sci.,
274, 1637 (2007).
[29] H. T. Diep, Statistical Physics: Fundamentals and Ap-
plication to Condensed Matter, Lectures, Problems and
Solutions (World Scientific, Singapore, 2015), chapter 3.
[30] R. J. Baxter, Exactly Solved Models in Statistical Physics,
(Academic Press, California, USA, 1982).
[13] J. K. Parrish, S. V. Viscido, and D. Grunbaum, Biol.
[31] J. Frohlich and T. Spencer, Communications in Mathe-
Bull. 202, 296 (2002).
[14] M. Adioui, J. P. Treuil, and O. Arino, Ecol. Modell. 167,
19 (2003).
[15] M. Zheng, Y. Kashimori, O. Hoshino, K. Fujita, and
T. Kambara, J. Theor. Biol. 235, 153 (2005).
[16] C. W. Reynolds, Flocks, herds, and schools: a distributed
behavioral model, Computer Graphics, SIGGRAPH'87
Conference Proceedings 21, 25 (1987).
[17] T. Vicsek, A. Czirok, E. Ben-Jacob,
I. Cohen,
matical Physics 81, 527-602 (1981).
[32] L. G´omez, A. Dobry, C. Geuting, H. T. Diep and L. Bu-
rakovsky, Phys. Rev. Lett. 90, 095701 (2003) and refer-
ences therein.
[33] D. R. Nelson and B. I. Halperin, Phys. Rev. B 19, 2457
(1979).
[34] W. Janke and H. Kleinert, Phys. Letters 105A, 134
(1984).
|
1206.2749 | 1 | 1206 | 2012-06-13T08:43:07 | Colour changes upon cooling of Lepidoptera scales containing photonic nanoarchitectures | [
"physics.bio-ph",
"cond-mat.mtrl-sci"
] | The effects produced by the condensation of water vapours from the ambient in the various intricate nanoarchitectures occurring in the wing scales of several Lepidoptera species were investigated by controlled cooling (from room temperature to -5 - -10 {\deg}C) combined with in situ measurement of changes in the reflectance spectra. It was determined that, due to this procedure, all photonic nanoarchitectures giving a reflectance maximum in the visible range and having an open nanostructure exhibited alteration of the position of the reflectance maximum associated with the photonic nanoarchitectures. The photonic nanoarchitectures with a closed structure exhibited little to no alteration in colour. Similarly, control specimens coloured by pigments did not exhibit a colour change under the same conditions. Hence, this effect can be used to identify species with open photonic nanoarchitectures in their scales. For certain species, an almost complete disappearance of the reflectance maximum was found. All specimens recovered their original colours following warming and drying. Cooling experiments using thin copper wires demonstrated that colour alterations could be limited to a millimetre, or below. Dried museum specimens do not exhibit colour changes when cooled in the absence of a heat sink due to the low heat capacity of the wings. | physics.bio-ph | physics | Colour changes upon cooling of Lepidoptera scales containing
photonic nanoarchitectures
István Tamáska1a, Krisztián Kertész1b, Zofia Vértesy1c, Zsolt Bálint2d, András Kun2e, Shen-
Horn Yen3f, László Péter Biró1g*
1Institute for Technical Physics and Materials Science, Research Centre for Natural
Sciences, P.O.B. 49, H-1525 Budapest, Hungary
2Hungarian Natural History Museum, Baross utca 13, H-1088 Budapest, Hungary
3Laboratory of Natural Resource Conservation, Department of Biology and Institute of Life
Science, National Sun Yat-Sen University, Kaohsiung, Taiwan, R. O. C.
Abstract
The effects produced by the condensation of water vapours from the ambient in the various
intricate nanoarchitectures occurring in the wing scales of several Lepidoptera species were
investigated by controlled cooling (from room temperature to -5 – -10 °C) combined with
in situ measurement of changes in the reflectance spectra. It was determined that, due to
this procedure, all photonic nanoarchitectures giving a reflectance maximum in the visible
range and having an open nanostructure exhibited alteration of the position of the
reflectance maximum associated with the photonic nanoarchitectures. The photonic
nanoarchitectures with a closed structure exhibited little to no alteration in colour.
Similarly, control specimens coloured by pigments did not exhibit a colour change under
the same conditions. Hence, this effect can be used to identify species with open photonic
nanoarchitectures in their scales. For certain species, an almost complete disappearance of
the reflectance maximum was found. All specimens recovered their original colours
following warming and drying. Cooling experiments using thin copper wires demonstrated
that colour alterations could be limited to a millimetre, or below. Dried museum specimens
do not exhibit colour changes when cooled in the absence of a heat sink due to the low heat
capacity of the wings.
Keywords: photonic nanoarchitectures; condensation; local cooling; display
Correspondence: [email protected], [email protected],
[email protected], [email protected], [email protected],
[email protected], [email protected], *Corresponding author
Introduction
Animal colour is mostly due to the spectrally selective reflection of incident light. It can be
associated with absorbing pigments (Zollinger et al. 2003; Nijhout 1991), which are often
referred to as chemical colour, and/or selective reflection caused by structural properties,
which is usually referred to as physical or structural colour (Joannopoulos et al. 2008; Biró
and Vigneron 2011). Some physical colours can be produced by photonic crystals (Seago et
al. 2009; Wilts et al. 2009; Shawkey et al. 2009a; Poladian et al. 2009; Ghiradella and
Butler 2009), which are periodic dielectric structures in space composed of two or more
media with different optical properties. These structures have a forbidden gap, within which
photons with certain energies cannot propagate and are completely reflected by the surface
of the photonic crystal. Since 1987, when Yablonovitch (Yablonovitch 1987) and John
(John 1987) described structures (i.e., nanoarchitectures) with photonic band gaps, the field
has been developing at an increasingly rapid pace.
An enormous variety of photonic structures can be found in nature. Butterflies, beetles
(Srinivasarao 1999, Seago et al. 2009, Shawkey et al. 2009a, Ghiradella & Butler 2009,
Biró et al. 2010) and even plants (Vigneron et al. 2005a, Glover & Whitney 2010) exhibit
these kinds of structures. It has recently been shown that, in certain species, such as the
sexually dichromatic species Hypolimnas bolina, females prefer conspecific males that
possess bright iridescent blue/ultraviolet dorsal ornamentation (Kemp 2007) so that the
"quality" of the photonic nanoarchitectures is well preserved from generation to generation.
As emphasised by several recent reports, visual signals from conspecific individuals make
important contributions to the mating behaviours of butterflies (Fordyce et al. 2002, Oliver
et al. 2002, Bálint et al. in press).
Several recent reviews have been published on photonic nanostructures of biological origin
(Parker 2002, Vukusic & Sambles 2003, Fratzl 2007, Kinoshita 2008, Biró & Vigneron
2011). Butterflies are diverse and highly practical examples of these structures, as their
colour is generated by pigments and by various nanostructures (Kertész et al. 2006a,
Kinoshita & Yoshioka 2005, Vértesy et al. 2006, Welch & Vigneron 2007, Vukusic et al.
1999, Vukusic & Stavenga 2009, Argyros et al. 2002). The essentially flat wings of
butterflies can be easily processed into convenient samples for both scientific examination
and practical applications, such as solar cells (Zhang et al. 2009), sensors (Potyrailo et al.
2007, Biró et al. 2008) and many other uses (Huang et al. 2006, Berthier et al. 2007).
Lepidopteran wings and the scales covering them have a complex geometric structure on
macroscopic- and microscopic-length scales. The membranes of the wings are generally
covered with partially overlapping scales on both sides. Typical scale dimensions are in the
ranges of 50 – 100 m for length and 15 – 50 m for width. Most species have two distinct
layers of different scales: ground scales and cover scales. While the underside of the scale
is rather featureless, the externally visible top surface usually exhibits a complex structure
from the micron to the nanometre range. This top surface or the volume of the scale
contains photonic nanostructures if the colour of the wings is of structural origin. These
nanostructures are mainly constructed of a chitinous matrix, including air holes (Scoble
1995, Kristensen 2003, Berthier 2006). These biological structures attract more and more
attention because they constitute a transition between random and crystalline order
(Shawkey et al. 2009; Liu et al. 2011). Recent papers reported that even a complete
photonic band gap can be found in such partially ordered materials (Florescu et al. 2009).
It often may not be a straightforward task to determine if a certain butterfly exhibits
structural colour, especially in the case of species presenting non-iridescent, matte
structural colour, as is seen on the ventral side of Cyanophrys remus (Kertész et al. 2006b)
or Callophrys rubi (Ghiradella & Radigan 1976). Of course, transmission and scanning
electron microscopy (TEM and SEM, respectively) can reveal structural details, but these
methods require time-consuming sample preparation (Shawkey et al. 2009b), expensive
equipment and physical destruction of the (potentially rare) exemplars. Sophisticated
optical characterisation (Kertész et al. 2006b, Wilts et al. 2009) can also reveal the presence
of photonic nanoarchitectures, but the required instruments and the expertise may not be
available in all laboratories; again, destructive sample preparation steps usually cannot be
avoided. The method described by Mason (Mason 1923), based on soaking the specimen
(sometimes for days) in liquids with a refractive index matching the refractive index of the
nanoarchitecture, may also be an alternative. Dropping liquids (e.g., oil or ethanol) onto
wings could identify nanostructures, but because these liquids cannot be found in the
natural environments of butterflies, this method may raise concern from museum curators
that it could damage the valuable samples. This type of method may induce persistent
changes in the status of a specimen.
In this paper, we present a simple method based on cooling in ambient air to observe and
investigate the colour-change of the wings of several Lepidoptera species coloured by
photonic nanoarchitectures in detail. Cooling the butterflies and allowing water vapour to
condense quickly onto the micro- and nanostructures induces colour change in their scales,
which possess structural colour arising from open nanoarchitectures. This simple method
allows the observer to determine which colours are of structural origin, as pigment-
coloured wings do not show the colour change observed in the case of structural colours.
During the cooling experiments, we identified significant differences in a few species in the
time lapse of the colour change and how it changed; these differences originate from
differences in air circulation in different nanoarchitectures. We also attempted to highlight
some basic ideas that underlie the interaction of water with wings and the colour-generating
nanoarchitectures. As softening in water vapour is the standard procedure for setting
butterflies, water vapour condensation should not raise concerns from museum curators. It
is well known that most Lepidoptera wings are hydrophobic (Wagner et al. 1996), which
prevents liquid water from penetrating into the photonic nanostructure because water
droplets roll off the surface (Wagner et al. 1996), thus preventing the soaking of the wings
in rain. Using the cooling method described herein, this difficulty can be easily avoided.
Materials and Methods
Eterusia taiwana (Lepidoptera: Zygaenidae) moths were obtained from the collections of
the Laboratory of Natural Resource Conservation, Department of Biology and Institute of
Life Science, National Sun Yat-Sen University, Taiwan. All other samples were obtained
from the Lepidoptera collection of the Hungarian Natural History Museum.
For structural investigation of the wings of Lepidoptera species, scanning electron
microscopy (SEM) and transmission electron microscopy (TEM) were used. SEM samples
were prepared by cutting off pieces of the wings, which were then attached to stubs with
double-sided conductive tape. All samples were subsequently coated with 15 nm sputtered
gold to allow for examination without charging. Cross-sectional TEM samples were
prepared by embedding pieces of wings in special resin. Thin sections with a thickness of
70 nm were cut using an ultramicrotome and were transferred to copper grids.
For optical characterisation, reflectance measurements were performed with an Avaspec
2048/2 fibre-optic spectrometer. A coaxial illuminator/pick-up fibre was used under normal
incidence conditions. Measurements were performed with unpolarised light using a
combined halogen-deuterium light source. An Avaspec diffuse white standard was used as
a reference sample. This system allows accurate measurements in the 200 – 1000 nm
wavelength range.
Measurements performed under controlled cooling were combined with spectroscopic
observation of the colour change. Three types of cooling apparatuses were used: (i) a
custom-built Peltier cooler, which could be placed under the combined illuminator/pick-up
fibre of the spectrometer; (ii) the deep freezer compartment of a commercially available
refrigerator; and (iii) thin copper wires frozen into ice blocks within the deep freezer
compartment. Peltier elements use the thermoelectric effect to directly convert electric
power to a temperature difference. They are usually thin plates (3 mm in our experiments)
and have two large sides (squares with 5 cm sides in our experiments) where the
temperature difference can be created. Two Peltier elements in series were mounted on a
CPU cooler fan for better heat dissipation. A shield was mounted around the Peltier
elements to direct the airflow away from the samples and to ensure a uniform temperature.
With this construction, a temperature of -5 ºC could be achieved in a normal room
environment. Spectroscopic measurements were taken every second during the controlled
cooling of the samples. An entire cooling cycle was typically completed over a time range
of minutes. While Peltier coolers are not expected to be present in all laboratories working
with butterflies, refrigerators are common in most of them. We performed the experiments
using the deep freezer compartment, cooling the butterflies for approximately 15 minutes.
Results
Colour change in wings due to cooling
The initial motivation for our work was provided by the experimental observation of colour
changes in living E. taiwana moths. The moths were kept in a refrigerator at approximately
5 °C to extend their lifespan. The cooled living E. taiwana moths exhibit colour changes
from green to dark brown to almost black when removed from the refrigerator. After being
maintained at room temperature for a few hours, the moth regained its former colouration.
The experiment was repeated with dried museum samples, but no colour changes were
observed. When cooling the dried wing pieces that had been removed from the bodies using
the Peltier cooler, colour changes could be easily observed (Fig. 1). The experiment was
repeated under an optical microscope. Water condensation (small water drops) and freezing
(ice) could be clearly observed on the surface of the Peltier cooler.
Further experiments were performed with wing samples from other Lepidoptera species
with photonic nanostructures in their scales that were cooled with Peltier elements [from
Lycaenidae: Albulina metallica (blue dorsal wing surface), Callophrys rubi (green ventral
wing surface), Cyanophrys remus (blue dorsal and green ventral wing surfaces) and
Polyommatus daphnis (blue dorsal wing surface); and from Nymphalidae: Morpho aega
(blue dorsal wing surface)]. As the width of the Peltier element is 5 cm, only small pieces
of lepidopteran wings were used to allow simultaneous cooling of several samples. Colour
changes could be easily observed on every sample (Fig. 2).
A simpler cooling method involves the use of a household refrigerator (cooling
temperature: -10 oC). Whole specimens can be cooled with an individual heat sink (glass
microscope slides) mounted under each wing, allowing non-destructive measurements to be
performed. Care has to be taken to provide good thermal contact between the microscope
slide and the wing; a thin wire will suffice for this purpose, as shown in Fig. 3. Four
butterflies were mounted on a large substrate with glass plates under each wing. The
samples were cooled for 30 minutes in the refrigerator before being removed to an ambient
room environment. The necessity of the heat sinks can be observed on butterflies in Fig. 3
a. and c. Colour changes cannot be observed near their bodies, where the wings are rapidly
heated due to the absence of a cold thermal sink. As lepidopteran wings are not rigorously
flat, some parts of the wings that are slightly closer to the cooled surface exhibit an earlier
colour change than others. A sample of the butterfly Lycaena virgaureae was used for
comparison, as it is not coloured by photonic nanoarchitectures, but by pigment (Vigneron
et al. 2005b). It did not exhibit any colour change upon cooling.
Spectroscopic measurements
E. taiwana wing samples were mounted on the surface of the top Peltier element, and
spectroscopic measurements were performed in situ during cooling (Fig. 4). The
measurements began from a dry status (curve 1 in Fig. 4). After applying the previously
determined voltage to the Peltier elements, the temperature dropped below the dewpoint in
a few minutes (approx. 10 °C). When the dewpoint was reached and condensation of the
water vapour began, the reflectance maximum shifted relatively rapidly to longer
wavelengths (curve 2 in Fig. 4), and the green/blue moths became dark-brown/green. The
samples were maintained at that temperature range for approximately 5 minutes after the
shift, but the reflectance spectrum did not change. Then, the temperature was set to below 0
°C, and the wing pieces became completely frozen. After for waiting 5 minutes during
which the position of the reflectance maximum did not change (curve 3 in Fig. 4), the
applied voltage was turned off to allow the samples heat up to ambient temperature (23 °C).
When melting began, the reflectance maximum shifted to shorter wavelengths with rapid
progress (from curve 3 to curve 4 in Fig. 4) and stopped for approximately one minute
when the ice in the wings was completely melted (0 °C). In a similar process, when the
water in the samples began to evaporate, the reflectance maximum shifted towards shorter
wavelengths and increased further in intensity until a colour close to the original colour was
achieved. Full restoration of the wing colour was accomplished by drying overnight,
indicating that a certain small but measurable fraction of water penetrated into the chitinous
material. Similar spectroscopic measurements were performed on several other
lepidopteran species, such as the well-characterised species C. remus (Kertész et al. 2006b)
(Fig. 5). During cooling, this green species became brown. Additionally, the temperature of
the Peltier cooler and the shift of the reflection maximum can be seen for the green ventral
side of A. metallica in Fig. 1 of the Supplementary Material.
Further experiments were performed on several species to investigate the time dependence
of the colour change (see Fig. 2 of the Supplementary Material for an example). The
samples were mounted on the Peltier cooler and cooled to approximately 10 °C, the
temperature at which condensation can begin. After 300 – 400 seconds, the cooling
apparatus was turned off, and the samples were allowed to warm up to ambient
temperature. The time-dependent shift of the reflectance peaks was investigated. The
resulting curves depend on the complex structural differences between the wing scale
nanoarchitectures and vapour condensation rates. Further investigation with more precise
control of environmental conditions will be required to obtain additional insight into the
processes taking place.
Repeated cooling – heating cycles were used to examine the long-term behaviour of the
spectral shift. The samples were cooled (approx. 10 °C), and after the reflectance maximum
had not changed for 5 minutes, the samples were heated to ambient temperature (23 °C);
following an additional 5 minutes, the whole cycle was repeated. After approximately the
third cycle, the butterflies completely lost their colour. Furthermore, chitin absorbed water,
and the wings became malleable. Drying the samples for a few days fully restored the
initial conditions of the wings. Observation of wing softening is not surprising, as this is the
common preparation procedure for fixing the position of museum specimens.
Local cooling by heat sink
This cooling method can also be used to change the colour of lepidopteran wings locally.
Cooling wings in specified small areas makes it possible to “write” on the wing surface.
One method for this involves the use of a matrix built from miniature Peltier cells (each cell
is used as a pixel), but a simpler method involves using cold wires below the wings. A cold
wire cools a small area around it. If the temperature around the wire is lower than the
dewpoint, condensation begins and changes the colour of the wing.
In our experiments, one end of thin copper wires was frozen into ice blocks, while the other
end was used to form horizontal characters above the ice blocks. The butterfly C. rubi was
placed onto these characters, and the characters were shaped to provide the best thermal
contact with the wing membrane (Fig. 6). Copper has very good heat conductivity; it can
easily transfer heat from the wings to the ice blocks, keeping the area cool around the wires
on the wings. The temperature in the cooled area can be set by adjusting the distance
between the ice and the characters. If the temperature of the wires is sufficiently high to be
close to the dewpoint, the coloured area around them will be narrower. If the wing is too
close to the ice block, the temperature of the entire wing can be lower than the dewpoint;
this causes the entire wing to change colour. The coloured area around the wires widens in
time as water vapour condenses. The writing begins to appear a few seconds after placing
the butterfly onto the wires, and it can be clearly observed after waiting another minute.
Discussion
Role of the heat sink in the colour change
The difference between the dried and living E. taiwana moths in our first observations is
that the body of the living specimens has a higher heat capacity due to its water content,
and it is able to act as a heat sink. In a resting position, these moths keep their closed wings
near their bodies, unlike many butterflies. The dried museum specimens have their wings
positioned on the sides of their bodies, as is usual for museum exemplars. Additionally,
they have lost most of the water content present in the living insect. Thus, they have a low
heat capacity, so they cannot provide an adequate heat sink; the wings are heated to room
temperature before significant vapour condensation could take place. This might be the
reason that these specimens do not exhibit any colour change after being removed from the
deep freezer compartment of the refrigerator (cooling temperature: -10 °C). If a large heat
sink is provided by mounting microscope slides under the wings to increase the heat
capacity of the combined system (wing + microscope slide), the wings can be kept cool for
a sufficient amount of time to allow condensation to take place.
Investigation of the possible biological significance of the colour changes of wings (when
cooled below the dewpoint in an ambient atmosphere) with respect to the survival strategy
of these species might be a very interesting subject for visual ecology, but it is outside the
scope of this paper.
Water penetration and wing softening
As chitin absorbs water, it swells, thus altering the thickness of small structures made of
chitin. This swelling effect can be observed in Figs. 4 and 5, where the reflection curves do
not immediately return to the initial position, and the maxima have lower intensities after
the cooling-heating cycle. However, overnight drying will fully recover the initial shape
and position of the reflection curves. Because total swelling of the structures can be
achieved by several cooling cycles (usually more than three), it is important that the
reflection variations of several species are compared in the first cooling cycle to minimise
the swelling effect.
When preparing museum specimens, butterflies are usually softened in saturated water
vapours for hours or days to cause them to become malleable so they can be set. This is a
slow process, as water vapours have to penetrate into the volume of the specimen aided by
daily temperature variations. Although a sample absorbs significant amounts of water, in
most cases, no liquid water will be present on its surface. A relaxed status of the wings is
achieved from the repeated cooling-heating cycles within a few minutes instead of the days
required for saturated water vapours.
Colour change and intensity variations in spectroscopic measurements
The photonic crystal structures in lepidopteran wings are mainly constructed from chitin
and air. Chitin has a moderate refractive index (n =1.56) (Vukusic et al. 1999). The position
of the reflectance maximum (λmax) can be easily estimated with a simple equation:
λmax=neff*d/m (Vigneron et al. 2006), where neff is the average refractive index; d is the
thickness of one period in a periodic structure; and m is an integer. When air (n = 1) is
replaced with water (n = 1.33), the contrast of the refractive indices decreases, which in
turn will modify the average refractive index value. This decrease of the refractive index
contrast shifts the stop-band (i.e., the reflectance maximum) to higher wavelengths, as can
be observed in the spectral measurements presented in Figs. 4 and 5 between the curves
labelled 1 and 2 The reverse process can also be observed between the curves labelled 4
and 5 Freezing the samples generates quite different results. For C. remus, the position of
the reflectance maximum is changed only moderately by freezing, while for E. taiwana, a
clear spectral shift occurs. Structural differences are likely responsible for these different
behaviours, as will be discussed below in the context of the SEM figures.
The change in the intensity of the reflectance maximum during the cooling experiments
depends on various parameters, such as the modifications induced in the surface geometry
of the wing by wetting and freezing, the deformation of the chitin structures due to water
absorption and the presence of water and ice on top of the scales. All of these parameters
can differ from individual to individual. Due to this complex behaviour, a precise
interpretation of the observed variations in intensity is beyond the scope of this paper.
Effect of the nanoarchitecture during cooling
During the simultaneous cooling experiments with the Peltier cooler (Fig. 2), it was
observed that the wings of different Lepidoptera species exhibited slight differences in the
time elapsed from the initiation of cooling until colour changes were clearly observed. This
can be explained by taking into account the differences in the colour-generating
nanoarchitectures in the scales.
If the surface temperature drops below the dew point, condensation begins on the surface,
and it is quite difficult to calculate the exact condensation rate. The diffusion and heat
equations have to be solved in parallel, making the boundary conditions match (for
example, see Kandlikar 1999). We will not attempt to solve this problem, but we will
attempt to highlight some essential features of the process. Lepidoptera wings are quite
complex structures, possessing a wing membrane and usually several layers of scales.
Colour is measured in the upper part of the wing, and cooling is applied to the lower part.
The transportation of heat from the upper scales to the lower scales is quite complex.
Therefore, the interactions between different scales and between the scales and the
membrane have to be taken into account. As the observation of colour change occurs on the
upper part of the wing, we attempt to restrict our interpretation to this part. When water
vapour begins to condense, it will decrease the relative humidity below the critical value;
for further condensation, a new supply of wet air is required. The more open the structure
is, the more easily the supply of wet air will penetrate it. The fastest colour changes were
observed for C. remus, P. daphnis and C. rubi, which have similar nanoarchitectures, while
slightly slower colour modifications were observed for A. metallica and E. taiwana. The
scales of E. taiwana exhibit a quite different, more closed structure compared to the other
species examined in this work; the surfaces of their scales have fewer and smaller holes
than the scales of the investigated butterflies (Fig. 7). It is possible that in closed
nanoarchitectures, such as those of Chrysiridia ripheus, the supply of wet air will be
insufficient for colour changes to occur (Fig. 7d). The cooling experiments demonstrated
this to be the case; no significant colour change could be observed in this species.
Conclusions
Our investigation of the rapid cooling of butterfly and moth wings in ambient air
demonstrated that the wings of these species, which have photonic crystal-type
nanoarchitectures with open structures, exhibit clear changes in coloration. Our data show
that the durations and magnitudes of the colour changes and the ways in which they
occurred were specific for certain structures (i.e., species), but small variations may also
occur from individual to individual. In particular, in some closed photonic
nanoarchitectures, such as those of the moth C. ripheus, a rapid colour change is not
observed because water vapour cannot easily penetrate such closed structures. Therefore, if
colour change is observed on cooling, this is a clear indication that the scales of the
examined butterfly and moth species are coloured by photonic nanoarchitectures, while the
very useful in the rapid screening of species to determine whether it may be worth
absence of the colour change cannot be interpreted as unambiguous evidence that the scales
are coloured only by pigments. In these cases, detailed structural (i.e., SEM and TEM) and
spectral investigation may be required. Nevertheless, this quick and simple test can prove
performing the time-consuming and expensive high-resolution characterisation procedures.
At the same time, this screen provides information regarding the open (C. remus, C. rubi)
or more closed (E. taiwana, A. metallica, C. ripheus) characters of the structures generating
the structural colours.
Cooling can be achieved using the deep freezer compartment of a common refrigerator if a
large heat capacity heat sink is mounted in good thermal contact with the wings. This
makes large-scale application of the method possible. It is worth pointing out that well-
dried collection specimens will not exhibit a colour change without an appropriate heat sink
under their wings. Another important advantage of cooling in the refrigerator may be that it
is non-destructive; even rare exemplars could be examined in this way without risking
damage to the samples.
If repeated cooling and heating cycles are performed in succession, the wings can be
softened in minutes. This rapid softening is attributed to repeated water vapour
condensation taking place in the interior of the micro- and nanostructures of the scales and
of the wing membranes. This will render the wings malleable without destruction of the
exemplars or the need to keep the butterflies in saturated water vapours for several days.
If local cooling is provided, such as by thin copper wires, it is possible to “write” on wings
possessing structural colours. This colour change effect upon water vapour condensation
could be used to construct a tuneable photonic crystal-based flat panel display. The colour
change can be induced by the local condensation of water vapours, which are always
present in ambient air. Each pixel should be mounted on a miniature Peltier element, and
the temperature (and colour) could be set with active feedback. The similar process of
capillary condensation has been employed for optical switching in porous optical
superlattices (Barthelemy et al. 2007) and is used by the beetle Dynastes hercules to
reversibly switch colour depending on the relative humidity of the environment (Rassart et
al. 2008).
Acknowledgements
The work in Hungary was supported by OTKA grant PD83483. The collaboration of
Hungarian and Taiwanese scientists was made possible by a joint agreement between the
Hungarian Academy of Science and the NSC in Taiwan for mobility. K. Kertész gratefully
acknowledges financial support from the János Bolyai Research Scholarship of the
Hungarian Academy of Sciences.
References
Argyros A, Manos S, Mckenzie DR, Cox GC, Dwarte DM. 2002. Electron tomography and
computer visualisation of a three-dimensional `photonic' crystal in a butterfly wing-scale.
Micron 33: 483-487.
Bálint Z, Kertész K, Piszter G, Vértesy Z, Biró LP. In press. The well-tuned blues: The role
of structural colours as optical signals in the species recognition of a local butterfly fauna
(Lepidoptera: Lycaenidae: Polyommatinae), Journal of the Royal Society Interface.
Barthelemy P, Ghulinyan M, Gaburro Z, Toninelli C, Pavesi L, Wiersma DS. 2007. Optical
switching by capillary condensation. Nature Photonics 1: 172-175.
Berthier S. 2006. Iridescences: The Physical Colors of Insects, 1st edn. Springer.
Berthier S, Boulenguez J, Bálint Z. 2007 Multiscaled polarization effects in Suneve
coronata (Lepidoptera) and other insects: application to anti-counterfeiting of banknotes.
Applied Physics A: Materials Science & Processing 86: 123-130.
Biró LP, Kertész K, Vértesy Z, Bálint Z. 2008. Photonic nanoarchitectures occurring in
butterfly scales as selective gas/vapor sensors. In: Creath K, Editor. The Nature of Light:
Light in Nature II. Proc. of SPIE Vol. 7057: 705706
Biró LP, Kertész K, Horváth E, Márk GI, Molnár G, Vértesy Z, Tsai J-F, Kun A, Bálint Z,
Vigneron J-P. 2010. Bioinspired artificial photonic nanoarchitecture using the elytron of
the beetle trigonophorus rothschildi varians as a ‘blueprint’. Journal of the Royal Society
Interface 7: 887-894.
Biró LP & Vigneron JP. 2011. Photonic nanoarchitectures in butterflies and beetles:
valuable sources for bioinspiration. Laser & Photonics Review 5, No. 1: 27-51.
Florescu M, Torquato S & Steinhardt PJ. 2009. Designer disordered materials with large,
complete photonic band gaps. Proceedings of the National Academy of Sciences of the
United States of America 106: 20658-63.
Fordyce JA, Nice CC, Forister ML, Shapiro AM. 2002. The significance of wing pattern
diversity in the Lycaenidae: mate discrimination by two recently diverged species.
Journal of Evolutionary Biology 15: 871-879.
Fratzl P. 2007. Biomimetic materials research: what can we really learn from nature’s
structural materials? Journal of the Royal Society Interface 4: 637-642.
Ghiradella HT, Radigan W. 1976. Development of butterfly scales: II. Struts lattices and
surface tension. Journal of Morphology150: 279-298.
Ghiradella HT, Butler MW. 2009. Many variations on a few themes: a broader look at
development of iridescent scales (and feathers). Journal of the Royal Society Interface 6:
S243-S251.
Glover BJ and Whitney HM. 2010. Structural colour and iridescence in plants: the poorly
studied relations of pigment colour. Annals of Botany 105 (4): 505-511.
Huang J, Wang X, Wang ZL. 2006. Controlled Replication of Butterfly Wings for
Achieving Tunable Photonic Properties. Nano Letters 6 (10): 2325-2331.
Joannopoulos JD, Meade RD, Winn JN. 2008. Photonic Crystals: Molding the Flow of
Light, 2nd edition. Princeton University Press.
John S. 1987. Strong localization of photons in certain disordered dielectric superlattices.
Physical Review Letters 58: 2486-2489.
Kandlikar SG. 1999. Handbook of Phase Change: Boiling and Condensation. Taylor &
Francis.
Kemp DJ. 2007. Female butterflies prefer males bearing bright iridescent ornamentation.
Proceedings of the Royal Society B: Biological Sciences 274: 1043-1047.
Kertész K, Bálint Z, Vértesy Z, Márk GI, Lousse V, Vigneron JP, Biró LP. 2006a. Photonic
crystal type structures of biological origin: Structural and spectral characterization.
Current Applied Physics 6: 252-258.
Kertész K, Bálint Z, Vértesy Z, Márk GI, Lousse V, Vigneron JP, Rassart M, Biró LP.
2006b. Gleaming and dull surface textures from photonic-crystal-type nanostructures in
the butterfly Cyanophrys remus. Physical Review E 74: 021922.
Kinoshita S & Yoshioka S. 2005. Structural colors in nature: the role of regularity and
irregularity in the structure. Computer Physics Communications 6: 1442–1459.
Kinoshita S. 2008. Structural Color in the Realm of Nature. Singapore, World Scientific.
Kristensen NP. 2003. Lepidoptera, Moths and Butterflies: Morphology, Physiology and
Development, Vol. 2. In: Handbook of Zoology, Vol. 4, Part 36. Walter de Gruyter.
Liu F, Dong B, Zhao F, Hu X, Liu X & Zi J. 2011. Ultranegative angular dispersion of
diffraction in quasiordered biophotonic structures. Optics Express 19: 7750-5.
Mason CW. 1923. Structural Colors in Feathers. Journal of Physical Chemistry 27 (3):
201–251.
Nijhout HF. 1991. The development and evolution of butterfly wing patterns. Smithsonian
Institution Press
Oliver JC, Robertson KA, Monteiro A. 2002. The significance of wing pattern diversity in
the Lycaenidae: mate discrimination by two recently diverged species. Journal of
Evolutionary Biology 15: 2369-2375.
Parker AR. 2002. Natural Photonic Engineers. Materials Today 4: Sept. 26-31.
Poladian L, Wickham S, Lee K, Large MCJ. 2009. Iridescence from photonic crystals and
its supression in butterfly scales. Journal of the Royal Society Interface 6: S233-S242.
Potyrailo RA, Ghiradella H, Vertiatchikh A, Dovidenko K, Cournoyer JR & Olson E. 2007.
Morpho butterfly wing scales demonstrate highly selective vapour response. Nature
Photonics 1: 123-128.
Rassart M, Colomer JF, Tabarrant T and Vigneron JP. 2008. Diffractive hygrochromic
effect in the cuticle of the hercules beetle Dynastes Hercules. New Journal of Physics 10:
033014.
Scoble MJ. 1995. The Lepidoptera: form, function, and diversity, 2nd ednition. Oxford
University Press.
Seago AE, Brady P, Vigneron, JP, Schultz TD. 2009. Gold bugs and beyond: a review of
iridescence and structural colour mechanisms in beetles (Coleoptera). Journal of the
Royal Society Interface 6: S165-S184.
Shawkey MD, Morehouse NI, Vukusic P. 2009a. A protean palette: colour materials and
mixing in birds and butterflies. Journal of the Royal Society Interface 6: S221-S231.
Shawkey MD, Saranathan V, Palsdottir H, Crum J, Ellisman MH, Auer M, Prum RO.
2009b. Electron tomography, three-dimensional Fourier analysis and colour prediction of
a three-dimensional amorphous biophotonic nanostructure. Journal of the Royal Society
Interface 6: S213-S220.
Srinivasarao M. 1999. Nano-Optics in the Biological Word: Beetles, Butterflies, Birds and
Moths. Chemical Reviews 99: 1935-1961.
Vértesy Z, Bálint Z, Kertész K, Vigneron JP, Lousse V, Biró LP. 2006. Wing scale
microstructures and nanostructures in butterflies--natural photonic crystals. Journal Of
Microscopy 224: 108-10.
Vigneron JP, Rasart M, Vértesy Z, Kertész K, Sarazzin M, Biró LP, Ertz D, Lousse V.
2005a. Optical structure and function of the white filamentary hair covering the
edelweiss bracts. Physical Review E 71: 011906.
Vigneron JP, Lousse V, Biró LP, Vértesy Z, Bálint Z. 2005b. Reflectance of topologically
disordered photonic-crystal films. In: Adibi A, Lin SY, Scherer A, Editors. Photonic
crystal materials and devices III, vol. 5733. Proceedings of SPIE, pp. 308–315.
Bellingham, WA: SPIE.
Vigneron JP, Rassart M, Vandenbem C, Lousse V, Deparis O, Biró LP, Dedouaire D,
Cornet A, Defrance P. 2006. Spectral filtering of visible light by the cuticle of metallic
woodboring beetles and microfabrication of a matching bioinspired material. Physical
Review E 73: 041905-1.
Vukusic P, Sambles JR, Lawrence CR, Wootton RJ. 1999. Quantified interference and
diffraction in single Morpho butterfly scales. Proceedings of the Royal Society B:
Biological Sciences 266: 1403–1411.
Vukusic P, Sambles JR. 2003. Photonic structures in biology. Nature 424: 852-855.
Vukusic P, Stavenga DG. 2009. Physical methods for investigating structural colours in
biological systems. Journal of the Royal Society Interface 6: S133-S148.
Wagner T, Neinhuis C, Barthlott W. 1996. Wettability and Contaminability of Insect
Wings as a Function of Their Surface Sculptures. Acta Zoologica 77: 213-225.
Welch VL, Vigneron JP. 2007. Beyond butterflies – the diversity of biological photonic
crystals. Optical and Quantum Electronics 39: 295-303.
Wilts BD, Leertouwer HL, Stavenga DG. 2009.
Imaging scatterometry and
microspectrophotometry of lycaenid butterfly wing scales with perforated multilayers.
Journal of the Royal Society Interface 6: S185-S192.
Yablonovitch E. 1987. Inhibited Spontaneous Emission in Solid-State Physics and
Electronics. Physical Review Letters 20: 2059-2062.
Zhang W, Zhang D, Fan T, Gu J, Ding J, Wang H, Guo Q, and Ogawa H. 2009. Novel
Photoanode Structure Templated from Butterfly Wing Scales. Chemistry of Materials 21:
33-40.
Zollinger H. 2003. Color Chemistry: Syntheses, Properties and Applications of Organic
Dyes and Pigments, 3rd revised edition. Wiley, Wenheim.
URE CAPT
FIGU
TIONS
a taiwana du
of Eterusia
Figur
ue to water v
n the wings
r change on
re 1: Colour
vapour
hindwing su
hite dorsal h
ue-black-wh
cooling. Blu
caused by c
nd freezing c
densation an
cond
urface
and g
green-white
e dorsal fore
ewing surfac
ce. a) At ro
om tempera
ature (22 °C
C); b) after v
vapour
cond
densation on
nto the struc
ctures (10 °C
C); c) after
freezing of
the wings (
(-5 °C); d) a
after
all of the w
vaporated fr
water had ev
ature, when
om tempera
eating to roo
ing and rehe
cooli
rom the
wing
gs.
efore cooling
ra wings be
f Lepidopter
of pieces of
consisting o
ple display
re 2: A simp
Figur
g (left)
and a
after cooling
g (right). Co
olour chang
ges can be o
observed du
e to water v
vapour cond
densation
om: a) Albu
ulina
wings are fro
eces of the w
les. The pie
s of the scal
ic structures
the photoni
onto
meta
); c) Cyanop
aega (blue d
b) Morpho a
allica (blue d
phrys
dorsal wing)
g surface); b
dorsal wing
remu
us (green ve
; d) Polyom
mmatus daph
hnis (blue d
); e) Cyanop
phrys
entral wing)
dorsal wing)
g) Eterusia t
ys rubi (gree
f) Callophry
us (blue dor
remu
taiwana
sal wing); f
wing); and g
en ventral w
(blue
e dorsal win
ng).
es with phot
b), c) Specie
erator. a), b
n the refrige
optera speci
re 3: Lepido
Figur
ies cooled i
tonic
d d) a speci
struc
ctures in the
eir scales an
ies without
photonic str
ructures in i
its scales. C
Colour
ecessity of t
s, and the ne
nostructures
photonic nan
ecies with p
seen for spe
nges can be
chan
the heat
can be obse
Under
not change.
colour did n
where the c
of a) and c),
the bodies o
erved near t
sink
ident
not change
wn in d) did
ments show
ured by pig
tterfly colou
ions, the bu
tical conditi
e colour.
a) Ca
allophrys ru
ubi, b) Cyan
nophrys rem
mus, c) Poly
yommatus be
ellargus, d)
) Lycaena
virga
aureae.
Figur
s of the
re 4: Spectr
ral measurem
ments on th
he blue/dors
sal a) and gr
reen/ventral
l b) surfaces
a taiwana. T
oth Eterusia
tionary refl
how the stat
ed curves sh
The numbere
gs of the mo
wing
ectance
(10 °C); 3. f
y status (23
tages: 1. dry
us cooling st
es in variou
curve
°C); 2. con
ndensation (
freezing
(-5 °C
C); 4. melti
ing (0 °C); a
and 5. evapo
oration proc
cesses.
es of the
al (b) surfac
green/dorsa
tral (a) and g
he blue/vent
ments on th
ral measurem
re 5: Spectr
Figur
mus. The num
nophrys rem
tterfly Cyan
he stationary
ves show th
mbered curv
gs of the but
wing
y
(23 °C); 2.
. dry status
us stages: 1.
ves in variou
ctance curv
reflec
condensatio
on (10 °C);
3.
freez
zing (-5 °C)
; 4. melting
g (0 °C); and
d 5. evapora
ation proces
sses.
re 6: Callop
phrys rubi w
were used t
to form char
Figur
was placed o
onto cooled
d wires that
racters.
form charac
was used to
other end w
ce, and the
rozen into ic
wires was fr
end of the w
One
cters (c).
The c
egin. The im
rea, where t
rrounding ar
cold wires c
cool the sur
the condens
sation can b
mages
the sample
fter putting
2-3 seconds
were
e collected 2
(a) and 1 m
minute (b) af
s onto the w
wires.
except for f)
f), where
ght: TEM, e
eft: SEM; rig
tterflies. Le
es of the but
TEM image
re 7: SEM/T
Figur
ys remus (bl
Cyanophry
lue dorsal w
re 2 m. a)
scale bars ar
mages. All s
are SEM im
both
wing
ace); b) Call
lophrys rub
bi (green dor
surfa
rsal wing su
urface, TEM
M image is r
reprinted fro
om
phnis (blue
mmatus dap
6); c) Polyo
g surface); d
dorsal wing
adigan 1976
adella & Ra
Ghira
d)
face); e) Alb
heus (blue-g
ysiridia riph
Chry
bulina metal
llica (blue d
al wing surf
green ventra
dorsal
f) Eterusia t
taiwana (gre
wing
g surface); f
een dorsal w
wing surface
e).
Electronic supplementary material
Cooling on
Freezing on
Cooling and
freezing off
F.S1: Position of the reflection maximum (gray line with symbols) on the green ventral
side of A. metallica during one whole cooling cycle (i.e., cooled below the dewpoint, then
frozen and then the cooling apparatus was turned off). The temperature of the top surface of
the Peltier cooler next to the wing (black line) was measured during one whole cooling
cycle (i.e., condensation, freezing, melting and evaporation).
: Time dep
F.S2
ecies (M.
ee blue spe
xima in thre
lection max
in the refl
f the shifts
pendence o
s) and in a s
is, C. remus
a, P. daphni
aega
ctures (gree
hotonic struc
h closed ph
species with
en part of
C. rip
pheus). The
t 350 sec
e cooling ap
pparatus wa
as turned on
n at 50 sec, a
and it was t
turned off a
the three
(gray
y lines). Rel
latively fast
t shifts of th
he reflectio
n maxima c
could be ob
bserved for
h faster shi
n, and much
s turned on
paratus was
cooling app
after the c
butterflies
blue
ifts were
obser
otonic struc
h closed pho
species with
d off. The s
t was turned
rved after it
ctures exhib
bited very
swelling
slow
w and small
shifts in bo
th cases, wh
hich could b
be interpret
ted as being
g due to the
e (epicuticle
pper surface
s in the up
open holes
e and/or the
he structure
of th
e very few
e) of the
nano
ostructure (s
see Fig. 7d).
.
|
1805.11312 | 3 | 1805 | 2018-09-20T21:34:54 | A new method for protein structure reconstruction from NOESY distances | [
"physics.bio-ph",
"q-bio.BM"
] | Protein structure reconstruction from Nuclear Magnetic Resonance (NMR) experiments largely relies on computational algorithms. Recently, some effective low-rank matrix completion (MC) methods, such as ASD and ScaledASD, have been successfully applied to image processing, which inspires us to apply the methods to reconstruct protein structures. In this paper, we present an efficient method to determine protein structures based on experimental NMR NOESY distances. ScaledASD algorithm is used in the method with several post-procedures including chirality refinement, distance lower (upper) bound refinement, force field-based energy minimization (EM) and water refinement. By comparing several metrics in the conformation evaluation on our results with Protein Data Bank (PDB) structures, we conclude that our method is consistent with the popularly used methods. In particular, our results show higher validities in Procheck dihedral angles G-factor. Furthermore, we compare our calculation results with PDB structures by examining the structural similarity to X-ray crystallographic structures in a special dataset. The software and its MATLAB source codes are available in https://github.com/xubiaopeng/PRASD | physics.bio-ph | physics | A new method for protein structure reconstruction from NOESY
distances
Z. Li, S. Li, X. Wei, X. Peng*, Q. Zhao*
ABSTRACT
Protein structure reconstruction from Nuclear Magnetic Resonance
(NMR) experiments largely relies on computational algorithms. Recently, some
effective low-rank matrix completion (MC) methods, such as ASD and ScaledASD,
have been successfully applied to image processing, which inspires us to apply the
methods to reconstruct protein structures. In this paper, we present an efficient method
to determine protein structures based on experimental NMR NOESY distances.
ScaledASD algorithm is used in the method with several post-procedures including
chirality refinement, distance lower (upper) bound refinement, force field-based
energy minimization (EM) and water refinement. By comparing several metrics in the
conformation evaluation on our results with Protein Data Bank (PDB) structures, we
conclude that our method is consistent with the popularly used methods. In particular,
our results show higher validities in Procheck dihedral angles G-factor. Furthermore,
we compare our calculation results with PDB structures by examining the structural
similarity to X-ray crystallographic structures in a special dataset. The software and
its MATLAB source codes are available in https://github.com/xubiaopeng/PRASD.
KEYWORDS: Protein structure reconstruction; NOESY distance restraints; matrix
completion.
INTRODUCTION
The characterization of three-dimensional protein structures has been a topic of great
interest for decades since the knowledge of the protein structures is essential to
understand the protein functions, which could lead to further pharmaceutical and
medical applications. Currently, X-ray crystallography and nuclear magnetic
resonance (NMR) spectroscopy (1) are two major experimental methods for protein
structure determination. Unlike X-ray crystallography, NMR spectroscopy is not a
"microscope with atomic resolution", but rather provides a network of distance
measurements between spatially proximate hydrogen atoms (2); however,
the
introduction of NMR spectroscopy to characterize protein structure is a breakthrough
because the method enables the identification of the protein structure in the aqueous
solution, which is closer to the states of the proteins in cells. Since the first
NMR-determined protein structure was reported in 1985 (1), thousands of protein
structures have been measured by NMR spectroscopy, which are available in the
1
Protein Data Bank (PDB) (3).
The typical NMR structure determination approach includes the following steps:
peak picking from NMR spectra, chemical shift assignment (spectral assignment),
geometric restraint assignment, and structural calculation (4). An important parameter
in NMR experiments is Nuclear Overhauser Effect (NOE), which is used to generate
distance restraints because the intensity of the NOE signal depends on the inverse
sixth power of the internuclear distance (5). The NOE intensity between two atoms is
reflected in the Nuclear Overhauser Effect Spectroscopy (NOESY), whose
assignment relies on the knowledge of chemical shifts of nuclei. However, not all H-H
distances can be measured accurately in the NMR experiments because the peak
intensities corresponding to the long distances (larger than 5 Å) are too weak to be
distinguished from the noise level in the experiment. Thus, usually just a network of
short distances is available in the protein NMR structure measurement.
Since NMR measurements only provide implicit information about the protein
structure, complicated computational algorithms have to be used to determine the
protein structure in NMR experiment. Molecular dynamics (MD) incorporated with
simulated annealing (SA) algorithm is one of the most widely used methods in protein
NMR structure determination. MD (6) simulation is based on classical mechanics and
often used to study the folding pathway and structure prediction (7-9), while SA (10)
is a heuristic global optimization method to find the conformation at energy minimum.
The SA procedure is usually not very fast. The performance of SA in different
NP-hard problems was evaluated by Johnson et al. (11,12). MD/SA hybrid methods of
XPLOR (13,14), DYANA (15), CYANA (16) and ARIA (17) were also implemented.
The XPLOR is a method built on the MD simulation package CHARMM (18). The
method can be used to search the conformations in Cartesian coordinate space. In
contrast, DYANA and its improved version CYANA work in torsion angle space and
hence have better calculation speed. ARIA can work in either torsion angle space or
Cartesian coordinate space, but the algorithms are optimized for ambiguous distance
restraints and violation analysis.
Euclidean distance matrix completion (EDMC) method is a different protein
reconstruction method which mainly relies on the distance matrix obtained from NMR
measurements. Compared to the MD/SA method, EDMC method uses fewer
assumptions in energy function. The pioneer work of protein NMR by EDMC was
done by Braun et al. in 1981 (19). Other efficient ways, such as EMBED (20) and
DISGEO (21), were developed by Havel's group in 1983 and 1984, respectively. Later,
semidefinite programming (SDP) was used to solve the EDMC problems by using
Gram matrix instead of distance matrix, including DAFGAL proposed by Biswas et al.
(22), DISCO proposed by Leung and Toh (23) and SPROS by Alipanahi et al. (24).
Here we introduce a new matrix completion algorithm for protein structure
reconstruction from NOESY distance restraints. By comparing our reconstructed
results with PDB structures, we demonstrate that the method is valid and comparable
with existing methods.
The paper is organized as follows: In Section 2, we describe our method for the
NMR protein structure reconstruction, including the establishment of the initial matrix,
2
matrix completion, distance bound refinement, chirality refinements, energy
minimization (EM) optimization and water refinement. In Section 3, we test the
validity of the method by comparing our calculation results with the PDB structures
from several different aspects. Finally, the conclusion and future work are presented in
Section 4.
METHODS
Similar to the method for image recovery, the protein structure can be expressed as a
distance matrix in the protein structure reconstruction, where each element is the
distance of a pair of atoms. In this distance matrix, some elements can be known from
the chemical properties, such as covalent bonds, and the NOESY distance restraints.
On the other hand, additional refinements including the refinements on chirality and
energetic stability in solution have to be applied to the recovered distance matrix due
to the special properties of the protein structure. Based on the above considerations,
we render the ScaledASD matrix completion algorithm, a successful algorithm in
image processing, to determine the distance matrix from NOESY data, followed by
further procedures to refine the structure.
Initial distance matrix establishment
To determine the protein structure using the matrix completion, firstly we need to
establish the initial distance matrix. The nonzero elements of this matrix stand for the
known atomic distances. There are two types of known distances that can be directly
filled into the matrix: 1) the distances measured in the NMR experiment, i.e., the
distance restraints from NOESY data; 2) the distances between the coplanar atoms
and all the covalent bonds. Although the coplanar and covalent bond distances can be
varied with different residues, for a specific residue type, they can be considered as
constants since the fluctuations are very small (25, 26).
However, a study has shown that it is insufficient to determine the structure
uniquely with the above known distances (27). It was proposed in the study that the
lower bound of the sample number m
in matrix completion theory should
, where n and r are the dimension and the rank of the matrix,
satisfy
m Cnr
log
10
n
respectively. The value of constant C is a certain positive number that is not exactly
known. In our protein distance matrix completion, we can estimate this number by
trials and tests. In detail, starting from a distance matrix of a protein with known
structure, we sample Cnrlog10n elements uniformly and reconstruct
the distance
matrix using these samplings, where C is tested with different values from 0.5 to 5.
For each C value we run recovery procedure ten times and calculate the average
RMSD. The relation between the RMSD and C value is shown in Fig. S1 in
Supplemental Information, indicating that the protein structure (distance matrix) is
well recovered when C is larger than 2. We take the lower limit C = 2. For a typical
NMR protein with 76 residues, such as 1G6J, atom number n~1200 and rank r = 5 for
Euclidean Distance Matrix (EDM) (28), we need at least m~36000 to rebuild the
matrix uniquely. In contrast, the NOESY data from NMR has only 1291 distances.
3
Together with all the covalent bonds and coplanar distances, there are 13488 distances
in total, which is still far smaller to meet the matrix completion requirement.
To solve this problem, we add some distance elements estimated by the triangle
inequality (29). In detail, assuming that there are N atoms named A1, A2, A3…AN, the
boundary of an unknown distance AiAj can be obtained from the known distances by
the following conditions: Max(AiAk - AkAj ) ≤AiAj ≤ Min (AiAk + AkAj) (for all k),
where Ak includes all the atoms whose distances AiAk and AkAj are known. In this
paper, we just take the upper boundary as the corresponding element in the initial
distance matrix for simplicity. Once the distance AiAj is known in this way, we repeat
this procedure including AiAj as the known elements to determine other unknown
distances until all the distance matrix elements are determined. Finally, the initial
distance matrix is established by putting the following three types of distances
together as described in Fig. 1: all
the
covalent/coplanar distances and a sampled subset of the distances estimated from the
triangle inequality.
the experimental NOESY data, all
FIGURE 1 The process of establishing an initial distance matrix.
Low-rank matrix completion
Matrix Completion (MC) (28, 30) problems are usually solved by minimizing the
rank of matrices or minimizing the nuclear norm. However, both methods are time
consuming (27,31). Recently, effective methods of alternating steepest descent (ASD)
and scaled variant ScaledASD (32) are reported to solve the matrix completion
problem directly with specified rank of matrix rather than using the nuclear norm
minimization. These approaches were used to update the solutions by incorporating an
exact
line-search based on the factorization of the variable. ScaledASD is an
accelerated version of ASD algorithm (32). In this paper, we only focus on
ScaledASD. Supposing we have a matrix Z0 with some elements are known in
4
advance and the complete matrix Z can be factorized into two parts X and Y, i.e.,
Z=XY. The ScaledASD algorithm can find the proper factors X and Y so that the
corresponding elements in matrix Z (Z=XY) are closest to the pre-known elements in
Z0. The flow chart of ScaledASD algorithm is
shown in Fig. 2, where
f(X,Y)=1/2PΩ(Z0)-PΩ(XY)F
the matrix
PΩ(Z0)-PΩ(XY) and PΩ is a sampling operator feching the known elements. The
optimization loop stops when the solutions get converged, and we get the completed
matrix at the end.
the Frobenius norm of
stands
for
FIGURE 2
The flow chart of ScaledASD algorithm.
Once the distance matrix is recovered, we can easily reconstruct the coordinates from
the Gram matrix following the reference (33).
Distance bound refinement
In fact, the distance estimation by triangle inequality introduces certain errors into the
initial distance matrix, which leads to inaccuracy in the structure calculation.
Meanwhile, because the presence of the internal motions and chemical exchange may
diminish the strength of the NOE signal (2), the interatomic distances obtained by
NOE experiment tend to be larger than the actual distances. Therefore, we perform a
distance bound refinement on the raw structure obtained from matrix completion to
reduce the above errors. For this purpose, we take the refinement procedure used in
reference (24), i.e., using nonlinear unconstrained optimization method BFGS (34) to
minimize the following function
(
X
)
E
L
i j E
( , )
i j L
( , )
x
i
g x
i
2
x
j
e
ij
x
j
l
ij
2
U
i j U
( , )
f
x
i
x
j
u
ij
2
R
i
1
x
i
2
,
(1)
where
f
(
) max(0,
)
and
g
(
) min(0,
)
. The sets E, U, and L represent
equality constraints, upper bounds and lower bounds, respectively; and
5
L
U
E
,
,
are their relevant coefficients. We take the same setting for parameters
E ,
2
1
U ,
1L
and
R
X
(
X
)
(0)
25 (
(0)
)
R
0
L
1,
U
R
E
0
25
1
n
i
1
.
2
(0)
x
i
(2)
where
1
1
is a parameter.
Similar to the procedures in reference (24), we set the covalent bond lengths and
distances among atoms on peptide plane as the equality constraints and a fraction of
the sum of the van der Waals radii of atoms between non-bonded atom pairs as the
lower bound to the corresponding distances:
r
(
(3)
),
vdw
d
ij
r
i
vdw
j
where
[0,1]
. is typically around 0.85 (35). The values of the van der Waals
radii are given in Table 1 (36, 37). However, we have different upper bounds in our
algorithm: we include both the measured data from the NOE experiment and the
estimated values from the triangle inequality as the upper bounds. As a result, by
minimizing Eq.(1) starting from the rebuilt structure in Section 2.2, we actually make
a complement on the upper limit approximation in both NOESY distance restraints
and the triangle inequality estimation in the initial distance matrix establishment
stage.
TABLE 1
Atom
vdwr
( Å)
Van der Waals radii for different atoms
O
1.4
H
1.0
C
1.7
N
1.5
Chirality refinements
Chirality is a geometric property of some molecules with mirror symmetry. As shown
in Fig. 3, although molecules A and B have the same distance matrix, they are
structurally different. In fact, the distance matrix does not provide any information
about the global chirality, and hence the chirality refinement has to be performed
additionally after the structure has been reconstructed from the distance matrix. Here,
we use the same chirality refinement method as in reference (24). In detail, two types
of chirality are considered: 1) The chirality of each residue: the residues in biological
protein are mostly left handed (L-type) as shown in Fig.3 B. If the chirality of a
residue is right handed, we just simply exchange positions of groups COOH and NH2.
2) The chirality of the Ramachandran angle : Most of the values should be
negative. If the number of
0 , we flip the
structure by taking the opposite sign of all the x-components of the coordinates.
is larger than the number of
0
6
FIGURE 3
enantiomer and (B) stands for the correct one in amino acids.
Schematic diagrams of chiral molecular structures. (A) stands for the incorrect
EM optimization and water refinement
The protein structure has to be energetically stable in solution. For this requirement,
we perform a force field-based EM optimization followed by water refinement on our
reconstructed structure after the above geometrical refinements. The purpose of EM
optimization is to get rid of the non-physical bonds and angles , and to find the
configuration corresponding to the local minimum in energy landscape. The TIP3P
water model and the AMBER99SB-ILDN (38) force field-based energy functions are
used in the calculation. With the EM optimized structure, we carry out the process of
thin-layer water refinement (39) where a simple annealing simulation on energy is
used to obtain the energetically favored structure in solution.
RESULTS AND DISCUSSION
We randomly picked ten proteins with different sizes and topologies measured by
NMR in the PDB references as listed in Table 2. Table 2 also lists the most reliably
well-defined regions by checking the NMR experimental report for each protein in
PDB. For the input, we extract the NOESY distances from DOCR database in the
NMR Restraints Grid (40,41), which is just a well-parsed database based on the NMR
experimental data file in PDB. Then, we do a sampling from the distances estimated
from the triangle inequality as described in Sec 2.1 at the ratio of 50/n, where n is
the number of atoms. The sampled elements are input into the initial distance matrix.
Recalling that the rank of a Euclidean distance matrix consistent with a protein
structure is only 5, we notice that under current sampling ratio the lower bound
requirement for matrix completion is satisfied for any protein with the number of
atoms less than 10,000. The distance matrix of protein structure is recovered with
ScaledASD method, followed by the geometrical refinements, EM optimization and
finally the water refinement. We notice that the EM optimization could be force field
dependent. However,
--
AMBER99SB-ILDN (38), CHARMM36m (42) and OPLSAA (43), the reconstructed
structure is robust to the force field selection. In the Supplemental Information, Table
S1 shows the Cα RMSD in the well-defined region between structures obtained from
different force fields. Therefore, we always use AMBER99SB-ILDN force field in our
EM optimization stage in the reconstruction.
tested with three
different
as we
force
fields
7
Clearly, our reconstructed structure is not unique due to the random sampling in
establishing the initial distance matrix. For each protein, we repeat our reconstruction
20 times. It is found that the average Cα RMSD in well-defined region between two
arbitrary structures from different runs is in the range 0.1~0.5Å, depending on the
number of atoms, showing clear convergence. All our tests were completed on a
desktop computer with a 3.4GHz processor and 8GB RAM. The matrix completion
and structural refinements were performed with the software MATLAB 2016a; EM
optimization and the water refinement were performed with GROMACS (44) and
XPLOR (13), respectively. Depending on the sizes of the proteins in the test, the
computation time for all our calculations varies from 10 minutes to 1.5 hours.
Information about ten test proteins
Description
Topology
α+β
β
Atoms Residues Well-defined
region in PDB
1-70
8-87
1228
1114
76
80
β
1267
125
2-125
Software
DYANA
XPLOR
CYANA
α/β
1648
104
9-70; 77-81;
89-101
NMRDRAW;
NMRPIPE
TABLE 2
PDB
ID
1G6J
1B4R
2K62
Ubiquitin
PKD domain 1
Liver fatty
acid-binding
protein
Yeast
ribosomal
protein
UPF0339
protein
SO3888
2L3O Murine
1CN7
2K49
2GJY
interleukin-3
Tensin 1 PTB
Domain
2YT0
2K7H Stress-induced
proteinSAM22
Amyloid
beta
A4 protein
Apolipoprotein
E
2L7B
α+β
α/β
α+β
α+β
1823
118
3-111
α
1980
127
43-138
2196
144
2337
157
5-61; 67-87;
95-104; 111-137
2-157
AutoStructure;
CYANA
CYANA;
XPLOR
DYANA
XPLOR
2602
176
18-30; 61-176
CYANA
α
4792
307
3-181; 186-200;
209-281;287-296
CYANA
Evaluation and comparison on reconstructed proteins
RMSD on well-defined region and TM-score comparison
To evaluate the performance of our method, we calculate the Cα RMSD values
between our reconstruction results and the first model in their corresponding PDB
structures in the well-defined regions as listed in Table 2. There are eight out of ten
proteins with RMSD < 2 Å, indicating that our reconstruction is generally successful.
8
For the two proteins with RMSD > 2 Å, one is just 2.08 Å, while the other is 2.21 Å
but with more than 300 residues. Reference (45) has shown that RMSD is length
dependent, i.e., larger protein tends to have larger RMSD with the same structural
similarity. To properly compare the structural similarity for the protein pairs with
different lengths, we employ a length independent quantity TM-score (45). TM-score
is a number in the range (0,1], where 1 indicates a perfect match between two
structures and 0 indicates a complete mismatch. Usually, a score higher than 0.5
indicates that the two structures have the same fold in SCOP/CATH (46). TM-scores
between our calculations and PDB structures are listed in the third column in Table 3,
showing very high similarity between them.
Finally, we show the corresponding protein structures in Fig. 4, where the red
represents the first model structure of the PDB references, and the blue represents one
model of our calculated structure.
TABLE 3
The Cα RMSD in well-defined region and TM-scores for ten proteins. The
RMSD and TM scores are calculated between the reconstructed ensemble and the first
model of the corresponding PDB structure. The standard deviation denotes the uncertainty
of the reconstructed structures
PDB ID
1G6J
1B4R
2K62
1CN7
2K49
2L3O
2GJY
2K7H
2YT0
2L7B
RMSD
/well-defined (Å)
1.00±0.10
1.57±0.16
1.81±0.14
1.53±0.13
1.52±0.10
2.08±0.19
1.63±0.09
1.85±0.22
0.95±0.07
2.21±0.14
TM-score
0.87±0.01
0.85±0.02
0.86±0.01
0.79±0.02
0.86±0.02
0.68±0.01
0.80±0.01
0.91±0.02
0.78±0.01
0.88±0.01
9
FIGURE 4 Superimposition of original PDB structures (red) and calculated structures (blue).
Structure Validation using MPscore and PSVS
To further validate our reconstruction method, we utilize the Molprobity (47) and
protein structure validation software suite (PSVS) (48) to evaluate the whole proteins
reconstructed from our algorithm.
In Molprobity, there is a quantity MPscore for assessing the overall quality of the
prediction structure from the statistical viewpoint. Lower values indicate better
models. PSVS is another NMR standard validation tool to check both the geometric
knowledge-based validation and the fit between the structure and the experimental
data. For NMR structures,
there are five geometric validation scores that are
noteworthy: Molprobity clash-score (47), Procheck Phi-Psi and all dihedral angle
G-factor (49), Verify3D score (50) and ProsaII score (51). In PSVS, they are all
rescaled to their corresponding Z-scores. Usually a higher Z-score indicates a better
model and a positive Z-score indicates that the analyzed structure is better than the
typical high-resolution X-ray structure. PSVS report also gives the number of
violations and maximum violation distance compared to the NOE experimental result.
Here we normalize the violation number into violation percentage using the number
of total restraints for comparison among proteins. Table 4 lists all these quantities in
both our calculation results and the corresponding PDB entries for comparison. As
shown in Table 4, our model generally performs better on Procheck Phi-Psi and all
dihedral angle G-factor. On some other evaluation metrics such as ProsaII, MPscore
and Molprobity clash-score, our calculation results are comparable with the
corresponding PDB deposits. On violation percentage and the maximum violation
distance, our calculation results are slightly worse than the PDB deposits.
10
1G6J
2K62
2K49
2L3O
1B4R
1CN7
Ref
Cal
Ref
Cal
Ref
Cal
Ref
Cal
Ref
Cal
Ref
Cal
Ref
Cal
2K7H Ref
Cal
Ref
Cal
Ref
Cal
2GJY
2YT0
2L7B
3.53
2.70
4.39
3.53
3.46
3.43
2.80
2.80
2.11
2.56
4.29
4.04
3.71
3.29
2.95
3.08
2.98
3.23
3.82
4.13
-0.96
-0.16
-1.93
-1.77
-0.80
-0.96
0.16
-0.64
-2.41
-2.25
-3.85
-3.85
-1.44
-2.89
-0.32
0.00
-0.96
-1.28
-3.05
-3.21
0.79
0.87
-1.94
-1.82
-0.54
-0.70
0.17
-0.62
-1.41
-1.36
-1.32
-1.49
-0.79
-1.28
-0.12
-0.21
0.41
0.37
-0.95
-1.41
-1.49/-4.55
-0.59/-1.30
-4.45/-7.75
-3.03/-5.26
-2.52/-4.26
-1.38/-3.43
-0.87/-2.60
-0.79/-1.95
-0.87/-0.71
-0.55/-1.06
-0.75/-5.32
0.24/-2.90
-2.28/-4.79
-1.93/-3.43
-0.35/-0.59
-0.83/-2.07
-1.97/-3.43
-0.71/-1.42
-0.47/-3.90
0.63/-1.89
Viol./
Cons.
(%)
0.49
5.38
5.91
20.29
36.82
26.18
2.71
10.45
9.78
11.39
17.72
24.73
4.74
11.47
14.48
16.19
22.33
23.65
MolP-
robity
Clash-
score
-5.77
-2.27
-23.36
-6.16
-1.79
-8.05
0.45
-2.09
-0.73
-1.90
-11.70
-11.07
-4.01
-4.23
-11.53
-4.14
-2.67
-3.24
-3.98
-12.45
Max
dist.
viol.
(Å)
0.41
2.58
2.29
2.74
7.06
3.22
1.23
2.71
2.51
2.78
3.65
4.16
1.08
3.68
4.73
2.77
2.07
2.27
TABLE 4 MPscores, the Z-score and RMS of distance violation/constraint of test protein
structures
Protein ID
ProsaII
MP-
score
Verify
3D
Procheck
(phi-psi/all)
21.01
26.56
3.22
4.03
Comparison on the protein structure properties
In Sec 3.1, we have evaluated our results and compared them with the corresponding
PDB structures using the scores provided in protein validation tools including
Molprobity and PSVS. All of the results are in the overall evaluation scores. For most
proteins, the local geometrical properties are sometimes more important in their
biological functions. In this section, we evaluate our results on two local geometrical
properties on backbone: the secondary structure and the Ramachandran Plot. In the
following, we just take one arbitrary PDB structure 2K7H as an example. For the
other proteins, similar results are obtained and listed in Supplemental Information.
Secondary structure comparison
Protein secondary structure is the 3D form of local segments, which is maintained by
hydrogen bonds formed between carbonyl and amide groups on the skeleton. The
secondary structure reflects the stability of protein structure. Since α-helices and
β-strands are the most common secondary structures, we compared the regions of
α-helices and β-strands in our calculated structures with the secondary structures in
PDB references, respectively. The secondary structures are classified by DSSP
11
algorithm (52). We select 2K7H as an example and show the secondary structures
from our calculation results and the comparisons with the PDB structures in Fig. 5.
Clearly, the secondary structures in the calculation results are almost the same as in
PDB models except the very short segment 75-77, which is identified as loop in our
calculation result but helix in the PDB structure. The similar comparisons for other
proteins are shown in Fig. S2 in Supplemental Information.
FIGURE 5 The secondary structure regions of 2K7H structure calculated based on ScaledASD
and the PDB reference structure.
Ramachandran Plot comparison
Ramachandran Plot is a way to visualize energetically allowed regions for backbone
dihedral angles and of amino acid residues in protein structures (53). Usually
there are three regions in Ramachandran Plot: favorable regions, allowed regions and
, ) in favorable regions indicate that the
disallowed regions. The combinations of (
results are without steric clashes; allowed regions indicate that the results are allowed
even if the steric constraints are slightly relaxed; and the disallowed regions indicates
that the results involve steric hindrance. We calculated the percentages of favourable
and disallowed regions in the Ramachandran Plot of the calculated structures by
MolProbity (47). The results of our calculation and the PDB structures are both listed
in Table 5, where we can see that the percentages in favourable region of our
calculated structures are comparable to the reference structures. To further confirm
that the Ramachandran angles of our calculated structure are in the correct region, we
compared the Ramachandran plots as shown in Fig. 6, where we can see the clear
consistence between our calculation result and the PDB structure, except that there are
two outliers in our calculation result.
TABLE 5 The percentages of the favourable and disallowed regions in Ramachandran Plot
for the ten calculated protein structures as well as PDB references
Protein
ID
1G6J
1B4R
2K62
Favourable
(calculation)
94.66±2.30
81.80±3.85
87.80±1.94
Favourable
(reference)
92.23±2.40
83.14±1.87
87.69±2.46
Disallowed
(reference)
0.63±0.91
3.91±1.14
0.93±0.56
Disallowed
(calculation)
1.01±1.51
6.73±1.71
2.40±1.66
12
1CN7
2K49
2L3O
2GJY
2K7H
2YT0
2L7B
79.51±2.56
94.57±1.66
74.18±2.02
75.99±3.12
97.29±1.08
85.55±2.76
75.69±1.58
84.12±2.78
92.99±2.39
70.36±4.07
78.72±3.04
93.06±2.78
79.67±2.39
73.38±1.92
4.02±1.71
0.30±0.42
8.86±1.85
7.29±1.96
1.10±0.24
1.15±0.56
6.41±1.19
2.89±1.89
1.47±0.97
12.64±3.07
6.20±1.51
1.33±1.10
6.07±1.36
9.11±1.61
FIGURE 6 The Ramachandran Plot of 2K7H structures: (A) our calculation result; (B) the
PDB reference structure.
13
Validation using X-ray crystallographic structure
Finally, we validate our method using a special dataset, which has both NMR
structures and the corresponding X-ray crystallographic structures in PDB. The
dataset is set up as follows: Among the ten proteins that we used in Sec. 3.1, there are
four proteins with their corresponding X-ray crystallographic structures. In addition,
we take the PDB entries listed in references (54, 55), and remove the proteins whose
RMSD values are larger than 3Å between the PDB NMR structures and the
corresponding X-ray ones. In the end, we get a dataset with 31 proteins. We calculate
the Cα RMSD values in the well-defined region between our calculated structure and
X-ray crystallographic PDB structure. In the meantime, we also calculate the same Cα
RMSD but for the PDB NMR structure compared with the X-ray crystallographic
structure in PDB. The results are shown in Fig.7 A. For 14/31 proteins our structures
are equivalent or even closer to the X-ray crystallographic structures than PDB NMR
structures; for the rest, 9 of our structures are just slightly further (ΔRMSD<0.5Å) to
the X-ray crystallographic structures than PDB NMR structures; for proteins with
PDB code 1BVM and 1MPH, our reconstruction results are visibly worse than the
PDB reference structures (ΔRMSD>1Å). We also compared the TM-score relative to
the X-ray structure between our calculated structure and corresponding PDB reference
structure, as shown in Fig.7 B. Similar to RMSD, the TM-scores of our structure are
generally slightly smaller than the PDB structure, but all are larger than 0.65
indicating that the foldings are still the same as the X-ray structures. One possible
reason that our calculated structure is slightly further to the X-ray crystallographic
structure than the PDB reference structure could be due to the fact that we only
consider the NOESY distance restraints in the NMR measurement while the PDB
structures take all the possible NMR measurements into consideration during the
reconstruction, which is the drawback of our current reconstruction algorithm.
14
FIGURE 7
(A) Cα RMSD in well-defined region relative to the X-ray crystallographic structure
for both our calculation result (blue line) and PDB structure (red line). The bar indicates the
standard deviation of the NMR ensemble, i.e., the reconstruction uncertainty. (B) Same as panel
(A) but measured for TM-score.
CONCLUSION
We have presented a new NMR structure determination method using ScaleASD
matrix completion algorithm. The difficulties to reconstruct the NMR structure using
matrix completion are from the following two facts: 1) The number of the measured
NOESY distance restraints is far smaller than the required number
in matrix
completion condition. 2) The NOESY distances are usually larger than the actual
atomic distances. To solve the first problem, we utilize the triangle inequality to
generate the proximate values for some unknown distances. Using the samplings from
these distances together with the measured NOESY data, covalent bonds and the
coplanar inter-atomic distances, we set up the initial distance matrix that satisfies the
matrix completion condition. For the second problem as well as the errors introduced
by the triangle inequality estimation, we perform the distance boundaries refinement
as complement. In the end, several refinements and optimization are implemented to
obtain more reasonable structure. The software and the MATLAB source codes are
distributed at https://github.com/xubiaopeng/PRASD. The release is archived in
Zenodo: DOI: 10.5281/zenodo.1400047.
To validate our method, ten arbitrarily selected proteins from NOESY data have
been reconstructed and compared with their corresponding structures in PDB. The
comparison includes several metrics such as Cα atom RMSD in well-defined region,
TM-score, MPscore, Z-scores and violations reported in PSVS software, secondary
15
structures and Ramachandran plots. It has been shown that our results are comparable
to the structures in PDB, especially for Procheck dihedral angles G-factor, the results
from our method show higher validity than the PDB structures. We have also
validated our reconstructed structure using the RMSD and TM-score with the
corresponding PDB X-ray crystallographic structure with an even larger dataset. The
result shows that our method is valid even though not as good as the current PDB
NMR structures. We notice that our current reconstruction structure is only based on
NOESY distance restraints. In the next step, many other NMR measurements such as
Residual dipolar coupling (RDC) would be considered and integrated into our
algorithm to obtain better structures.
AUTHOR CONTRIBUTIONS
Z. L. developed the algorithm, performed the calculation, analyzed the data,
validated the method and made the manuscript draft. S. L calculated and analyzed the
data. X.W. helped in data analysis, method validation and revised the manuscript. X. P.
supervised the data analysis and validation, designed the software interface and
revised the manuscript. Q. Z. designed the project, supervised the calculation as well
as analysis, and revised the manuscript.
ACKNOWLEDGMENTS
The authors would like to thank Prof. Ge Molin for valuable discussions, Prof. Deng
Rongping for reading the manuscript and improving English. This work was
supported by the National Science Foundation (NSF) of China with the Grant
No.11675014. Additional support was provided by the Ministry of Science and
Technology of China (2013YQ030595-3).
16
REFERENCES
1. Williamson, M. P., T. F. Havel, and K. Wüthrich. 1985. Solution conformation of
protein inhibitor IIA from bull seminal plasma by 1H nuclear magnetic resonance
and distance geometry. J. Mol. Biol. 182:295 -- 315.
2. Güntert, P. 1998. Structure calculation of biological macromolecules from NMR
3. Berman, H. M., J. Westbrook, ... , P. E. Bourne. 2000. The protein data bank,
17
data. Q. Rev. Biophys. 31: 145 -- 237.
Nucleic acids research 28. 1: 235 -- 242.
Biophysics. J. 38: 129-143.
4. Güntert, P. 2009. Automated structure determination from NMR spectra, Eur.
5. Pitner, T. P., J. D. Glickson, and G. R. Marshal. 1974. Solvent exposure of specific
nuclei of angiotensin II determined by NMR solvent saturation method. Nature.,
250: 582-584.
6. Mccammon, J. A., B. R. Gelin, and M. Karplus. 1977. Dynamics of folded
proteins. Nature. 267: 585-590.
7. Davtyan, A., N. P. Schafer, …, G. A. Papoian. 2012. Awsem-md: coarse-grained
protein structure prediction using physical potentials and bioinformatically based
local structure biasing. J. Phys. Chem. B. 116: 8494-8503.
8. Czaplewski, C., S. Oldziej, …, H. A. Scheraga. 2004. Prediction of the structures
of proteins with the unres force field, including dynamic formation and breaking
of disulfide bonds. Protein Eng. Des. Sel. 17: 29-36.
9. Lindorfflarsen, K., S. Piana, …, D. E. Shaw. 2011. How fast-folding proteins fold.
10. Kirkpatrick, S., C. D. Gelatt, and M. P. Vecchi. 1983. Optimization by simulated
Science. 334: 517-520.
annealing. Science. 220: 671-680.
11. Johnson, D. S., C. R. Aragon, …, C. Schevon. 1989. Optimization by simulated
annealing: an experimental evaluation. Part I, graph partitioning. Oper. Res. 37:
865-892.
12. Johnson, D. S., C. R. Aragon, …, C. Schevon. 1991. Optimization by simulated
annealing: An experimental evaluation; part II, graph coloring and number
partitioning. Oper. Res. 39: 378-406.
13. Schwieters, C. D., J. J. Kuszewski, …, G. M. Clore. 2003. The Xplor-NIH NMR
molecular structure determination package. J. Magn. Reson. 160: 65-73.
14. Schwieters, C. D., J. J. Kuszewski, and G. M. Clore. 2006. Using Xplor-NIH for
NMR molecular structure determination. Prog. Nucl. Magn. Reson. Spectrosc. 48:
47-62.
15. Güntert, P., C. Mumenthaler, and K. Wüthrich. 1997. Torsion angle dynamics for
NMR structure calculation with the new program DYANA. J. Mol. Biol. 273:
283-298.
16. Güntert, P. 2004. Automated NMR structure calculation with CYANA. Methods
Mol. Biol. 278: 353-378.
17. Linge, J. P., S. I. O'Donoghue, and M. Nilges. 2001. Automated assignment of
ambiguous nuclear overhauser effects with ARIA. Method. Enzymol. 339: 71-90.
18. Brooks, B. R. 1983. CHARMM: A program for macromolecular energy,
minimization, and dynamics calculations. J. Comput. Chem. 4: 187-217.
19. Braun, W., C. Bösch, …, K. Wüthrich. 1981. Combined use of proton-proton
Overhauser enhancements and a distance geometry algorithm for determination
of polypeptide conformations. Application to micelle-bound glucagon. Biochim.
Biophys. Acta. 667: 377-396.
20. Havel, T. F., I. D. Kuntz, and G. M. Crippen. 1983. The theory and practice of
distance geometry. Bull. Math. Biol. 45: 665-720.
21. Havel, T., and K. Wüthrich. 1984. A distance geometry program for determining
the structures of small proteins and other macromolecules from nuclear magnetic
resonance measurements of intramolecular 1H-H proximities in solution. Bull.
Math. Biol. 46: 673-698.
22. Biswas, P., K. C. Toh, and Y. Ye. 2008. A distributed SDP approach for
largescale noisy anchor-free graph realization with applications to molecular
conformation. SIAM J. Sci. Comput. 30: 1251 -- 1277.
23. Leung, N. H., and K. C. Toh. 2009. An SDP-based divide-and-conquer algorithm
for large-scale noisy anchor-free graph realization. SIAM J. Sci. Comput. 31:
4351 -- 4372.
24. Alipanahi, B., N. Krislock, and M. Li. 2013. Determining protein structures from
NOESY distance constraints by semidifinite programming. J. Comput. Biol. 20:
296 -- 310.
25. Engh, R. A., and R. Huber. 1991. Accurate bond and ample parameters for X-ray
protein structure refinement. Acta Cryst. 47: 392-400.
26. Hooft, R. W. W., C. Sander, and G. Vriend. 1996. Verification of protein
structures: side-chain planarity. J. Appl. Cryst. 29: 714-716.
27. Candès, E. J., and T. Tao. 2010. The power of convex relaxation: near-optimal
matrix completion. IEEE Trans. Inf. Theory. 56: 2053 -- 2080.
28. Candès, E. J., and Y. Plan. 2011. Tight oracle inequalities for low-rank matrix
recovery from a minimal number of noisy random measurements. IEEE Trans.
Inf. Theory. 57: 2342 -- 2359.
29. Warshall, S. 1962. A theorem on boolean matrices. JACM. 9: 11 -- 12.
30. Candès, E. J., and B. Recht. 2009. Exact matrix completion via convex
optimization. Found. Comput. Math. 9: 717 -- 772.
31. Harvey, N. J. A., D. R. Karger, and S. Yekhanin. 2006. The complexity of matrix
completion. SODA'06 Proceedings of the seventeenth annual ACM-SIAM
symposium on discrete algorithms. 1103 -- 1111.
32. Tanner, J., and K. Wei. 2016. Low rank matrix completion by alternating steepest
descent methods. Appl. Comput. Harmon. Anal. 40: 417-429.
33. Crippen G. M., and T. F. Havel. 1978. Stable calculation of coordinates from
distance information. Acta Cryst. 34: 282-284.
34. Lewis, A. S., and M. L. Overton. 2013. Nonsmooth optimization via
quasi-Newton methods. Math. Program., 141: 135-163.
35. Cassioli, A., B. Bardiaux, …, T. E. Malliavin. 2015. An algorithm to enumerate
all possible protein conformations verifying a set of distance constraints. BMC
18
Bioinformatics. 16: 1-15.
36. Rocchia, W., E. Alexov, and B. Honig. 2001. Extending the applicability of the
constants and
nonlinear poisson-boltzmann equation: Multiple dielectric
multivalent ions. J. Phys. Chem. B. 105: 6507 -- 6514.
37. Honig, B., and A. Nicholls. 1995. Classical electrostatics in biology and chemistry.
Science. 268: 1144 -- 1149.
38. Lindorfflarsen, K., S. Piana, …, D. E. Shaw. 2010. Improved side-chain torsion
potentials for the Amber ff99SB protein force field. Proteins. 78: 1950-1958.
39. Linge, J. P., M. Habeck, …, M. Nilges. 2003. ARIA: automated NOE assignment
and NMR structure calculation. Bioinformatics. 19: 315-316.
40. Doreleijers, J. F., S. Mading, …, E. L. Ulrich. 2003. BioMagResBank database
with sets of experimental NMR constraints corresponding to the structures of
over 1400 biomolecules deposited in the protein data bank. J. Biomol. NMR. 26:
139-146.
41. Doreleijers, J. F., A. J. Nederveen, …, E. L. Ulrich. 2005. BioMagResBank
databases DOCR and FRED containing converted and filtered sets of
experimental NMR restraints and coordinates from over 500 protein PDB
structures. J. Biomol. NMR. 32: 1-12.
42. Huang, J., S. Rauscher, S, …, A. D. MacKerell. 2017. CHARMM36m: an
improved force field for folded and intrinsically disordered proteins. Nat.
Methods. 14: 71-73.
43. Quiñonero, D., S. Tomàs, …, P. M. Deyà. 2001. OPLS all-atom force field for
squaramides and squaric acid. Chem. Phys. Lett. 350: 331-338.
44. Berendsen, H. J. C., D. V. D. Spoel, and R. V. Drunen. 1995. GROMACS: a
message-passing parallel molecular dynamics implementation. Comput. Phys.
Commun. 91: 43-56.
45. Zhang, Y., and J. Skolnick. 2004. Scoring function for automated assessment of
protein structure template quality. Proteins. 57: 702-710.
46. Hadley, C. and D. T. Jones. 1999. A systematic comparison of protein structure
classifications: SCOP, CATH and FSSP. Structure. 7: 1099-1112.
47. Chen, V. B., A. W. Rd, …, D. C. Richardson. 2010. Molprobity: all-atom structure
validation for macromolecular crystallography. Acta. Crystallogr. D. Biol.
Crystallogr. 66: 12-21.
48. Bhattacharya, A., R. Tejero, and G. T. Montelione. 2007. Evaluating protein
structures determined by structural genomics consortia. Proteins. 66: 778-795.
49. Laskowski, R. A., M. W. Macarthur, …, J. M. Thornton. 1993. Procheck -- a
program to check the stereochemical quality of protein structure. J. Appl.
Crystallogr. 26: 283-291.
50. Eisenberg, D., R. Luthy, and J. U. Bowie. 1997. VERIFY3D: assessment of
protein models with three-dimensional profiles. Method. Enzymol. 277: 396-404.
51. Wiederstein, M., and M. J. Sippl. 2007. ProSA-web: interactive web service for
the recognition of errors in three-dimensional structures of proteins. Nucleic.
Acids. Res. 35: W407-W410.
52. Kabsch, W., and C. Sander. 1983. Dictionary of protein secondary structure:
19
Pattern recognition of hydrogen-bonded and geometrical features. Biopolymers,
22: 2577-2637.
53. Ramachandran, G. N., C. Ramakrishnan,
and V. Sasisekharan. 1963.
Stereochemistry of polypeptide chain configurations. J. Mol. Biol. 7: 95-99.
54. Sikic, K., S. Tomic, and O. Carugo. 2010. Systematic comparison of crystal and
NMR protein structure deposited in the Protein Data Bank. Open Biochem. J. 4:
83-95.
55. Schneider, M., X. Fu, and A. E. Keating. 2009. X-ray vs. NMR structures as
templates for computational protein design. Proteins. 77: 97-110.
20
|
1110.6927 | 1 | 1110 | 2011-10-31T19:56:09 | Statistical-Mechanical Measure of Stochastic Spiking Coherence in A Population of Inhibitory Subthreshold Neurons | [
"physics.bio-ph",
"q-bio.NC"
] | By varying the noise intensity, we study stochastic spiking coherence (i.e., collective coherence between noise-induced neural spikings) in an inhibitory population of subthreshold neurons (which cannot fire spontaneously without noise). This stochastic spiking coherence may be well visualized in the raster plot of neural spikes. For a coherent case, partially-occupied "stripes" (composed of spikes and indicating collective coherence) are formed in the raster plot. This partial occupation occurs due to "stochastic spike skipping" which is well shown in the multi-peaked interspike interval histogram. The main purpose of our work is to quantitatively measure the degree of stochastic spiking coherence seen in the raster plot. We introduce a new spike-based coherence measure $M_s$ by considering the occupation pattern and the pacing pattern of spikes in the stripes. In particular, the pacing degree between spikes is determined in a statistical-mechanical way by quantifying the average contribution of (microscopic) individual spikes to the (macroscopic) ensemble-averaged global potential. This "statistical-mechanical" measure $M_s$ is in contrast to the conventional measures such as the "thermodynamic" order parameter (which concerns the time-averaged fluctuations of the macroscopic global potential), the "microscopic" correlation-based measure (based on the cross-correlation between the microscopic individual potentials), and the measures of precise spike timing (based on the peri-stimulus time histogram). In terms of $M_s$, we quantitatively characterize the stochastic spiking coherence, and find that $M_s$ reflects the degree of collective spiking coherence seen in the raster plot very well. Hence, the "statistical-mechanical" spike-based measure $M_s$ may be used usefully to quantify the degree of stochastic spiking coherence in a statistical-mechanical way. | physics.bio-ph | physics | Noname manuscript No.
(will be inserted by the editor)
Statistical-Mechanical Measure of Stochastic Spiking
Coherence in A Population of Inhibitory Subthreshold
Neurons
Woochang Lim · Sang-Yoon Kim
Received: date / Accepted: date
Abstract By varying the noise intensity, we study stochastic spiking coherence (i.e., col-
lective coherence between noise-induced neural spikings) in an inhibitory population of
subthreshold neurons (which cannot fire spontaneously without noise). This stochastic spik-
ing coherence may be well visualized in the raster plot of neural spikes. For a coherent case,
partially-occupied "stripes" (composed of spikes and indicating collective coherence) are
formed in the raster plot. This partial occupation occurs due to "stochastic spike skipping"
which is well shown in the multi-peaked interspike interval histogram. The main purpose
of our work is to quantitatively measure the degree of stochastic spiking coherence seen in
the raster plot. We introduce a new spike-based coherence measure Ms by considering the
occupation pattern and the pacing pattern of spikes in the stripes. In particular, the pacing
degree between spikes is determined in a statistical-mechanical way by quantifying the aver-
age contribution of (microscopic) individual spikes to the (macroscopic) ensemble-averaged
global potential. This "statistical-mechanical" measure Ms is in contrast to the conventional
measures such as the "thermodynamic" order parameter (which concerns the time-averaged
fluctuations of the macroscopic global potential), the "microscopic" correlation-based mea-
sure (based on the cross-correlation between the microscopic individual potentials), and the
measures of precise spike timing (based on the peri-stimulus time histogram). In terms of
Ms, we quantitatively characterize the stochastic spiking coherence, and find that Ms re-
flects the degree of collective spiking coherence seen in the raster plot very well. Hence,
the "statistical-mechanical" spike-based measure Ms may be used usefully to quantify the
degree of stochastic spiking coherence in a statistical-mechanical way.
Keywords Inhibitory Subthreshold Neurons · Stochastic Spiking Coherence · Statistical-
Mechanical Measure
1 Introduction
Recently, brain rhythms have attracted much attention (Buzs´aki 2006). Synchronous oscil-
lations in neural systems may be used for efficient sensory processing (e.g., binding of the
Woochang Lim · Sang-Yoon Kim (corresponding author)
Department of Physics, Kangwon National University, Chunchon, Kangwon-Do 200-701, Korea
E-mail: [email protected]
1
1
0
2
t
c
O
1
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
2
9
6
.
0
1
1
1
:
v
i
X
r
a
2
integrated whole image in the visual cortex is accomplished via synchronization of neural
firings) (Gray 1994). In addition to such neural encoding of sensory stimuli, neural syn-
chronization is also correlated with pathological rhythms associated with neural diseases
(e.g., epileptic seizures and tremors in the Parkinson's disease) (Hauschildt et al. 2006;
Grosse et al. 2002; Tass et al. 1998). Here, we are interested in these synchronized neu-
ral oscillations.
A neural circuit in the major parts of the brain such as thalamus, hippocampus, and
cortex consists of a few types of excitatory principal cells and diverse types of inhibitory in-
terneurons. Functional diversity of interneurons increases the computational power of prin-
cipal cells (Buzs´aki 2006; Buzs´aki et al. 2004). To understand the mechanisms of syn-
chronous brain rhythms, neural systems composed of excitatory neurons and/or inhibitory
neurons have been investigated, and thus three types of synchronization mechanisms have
been found (Wang 2003): recurrent excitation between principal cells, mutual inhibition
between interneurons, and feedback between excitatory and inhibitory neurons. Perfect syn-
chronization occurs in the network of pulse-coupled excitatory neurons (Kuramoto 1991;
Mirollo & Strogatz 1990). However, this synchronization via mutual excitation cannot be
always stable in networks with slowly decaying synaptic couplings. Stability of synchro-
nization is shown to depend on both the time course of synaptic interaction and the response
of neurons to small depolarization (Vreeswijk et al. 1994; Hansel et al. 1995). When the
decay time of the synaptic interaction is enough long, mutual inhibition (rather than exci-
tation) may synchronize neural firing activities. By providing a coherent oscillatory output
to the principal cells, interneuronal networks play the role of the backbones (i.e., synchro-
nizers or pacemakers) of many brain oscillations such as the 10-Hz thalamocortical spin-
dle rhythms (Wang & Rinzel 1992; Golomb & Rinzel 1994) and the 40-Hz fast gamma
rhythms in the hippocampus and the neocortex (Wang & Buzs´aki 1996; White et al. 1998;
Whittington et al. 2000; Tiesinga et al. 2001). When the feedback between the excitatory
and inhibitory populations is strong, neural synchrony appears via the "cross-talk" between
the two populations (Whittington et al. 2000; Tiesinga et al. 2001; Hansel & Mato 2003;
Borgers & Kopell 2003, 2005).
Most past studies exploring mechanisms of neural synchronization were done in neural
systems composed of spontaneously firing (i.e., self-oscillating) suprathreshold neurons. For
this case, neural coherence occurs via cooperation of regular firings of suprathreshold self-
firing neurons. In contrast, neural systems composed of subthreshold neurons have received
little attention. Unlike the suprathreshold case, each subthreshold neuron in the absence of
coupling cannot fire spontaneously without noise; it can fire only with the help of noise.
Recently, stochastic spiking coherence (i.e., collective coherence emerging via cooperation
of noise-induced spikings) was observed in an excitatory population of pulse-coupled sub-
threshold neurons (Wang et al. 2000; Lim & Kim 2007, 2009). This kind of works may be
thought to correspond to a "subthreshold version" of neural synchronization through mutual
excitation. Due to the stochastic spiking coherence, synaptic current, injected into each in-
dividual neuron, becomes temporally coherent. Hence, temporal coherence resonance of an
individual subthreshold neuron in the network may be enhanced. This enhancement of co-
herence resonance in an excitatory network of subthreshold Hodgkin-Huxley neurons was
characterized in terms of the coherence factor b , representing the degree of sharpness of
the peak in the power spectrum of an individual neuron (Wang et al. 2000). In this way, the
measure b
is used for characterization of the temporal coherence of an individual neuron.
However, we note that b
is not a measure for directly measuring the degree of collective
coherence in the whole population.
3
In this paper, we are concerned about the "subthreshold version" of neural synchro-
nization via mutual inhibition. In Section 2, we describe the biological conductance-based
Morris-Lecar (ML) neuron model with voltage-gated ion channels (Morris & Lecar 1981;
Rinzel & Ermentrout 1998; Tsumoto et al. 2006). The ML neurons (used in our study)
exhibit the type-II excitability (i.e., the firing frequency begins to increase from a non-
zero value when the stimulus exceeds a threshold value), and they interact via inhibitory
GABAergic synapses whose activity increases fast and decays slowly. In Section 3, we
characterize stochastic spiking coherence in a large population of inhibitory subthreshold
ML neurons by varying the noise intensity for a fixed coupling strength. Weakly coherent
states with oscillating ensemble-averaged global membrane potential VG are thus found to
appear in a range of intermediate noise intensity [i.e., regular global oscillation (with re-
duced amplitude and increased frequency) emerges via cooperation of irregular individual
oscillations]. Emergence of collective coherence may be well described in terms of the con-
ventional "thermodynamic" order parameter which concerns the time-averaged fluctuations
of the macroscopic global potential VG. We note that this stochastic spiking coherence may
be well visualized in the raster plot of neural spikes (i.e., a spatiotemporal plot of neural
spikes) which is directly obtained in experiments. For the coherent case, "stripes" (com-
posed of spikes and indicating collective coherence) are found to be formed in the raster
plot. Due to coherent contribution of spikes, local maxima of the global potential VG appear
at the centers of stripes. However, these stripes are partially occupied. Individual inhibitory
neurons exhibit intermittent spikings phase-locked to VG at random multiples of the period of
VG. This "stochastic phase locking" leading to "stochastic spike skipping" is well shown in
the interspike interval (ISI) histogram with multiple peaks. The multi-peaked ISI histogram
shows some indication of weak collective spiking coherence.
The main purpose of our work is to quantitatively measure the degree of stochastic spik-
ing coherence seen in the raster plot. In Section 3, we introduce a new type of spike-based
coherence measure Ms by taking into consideration the occupation pattern and the pacing
pattern of spikes in the stripes of the raster plot. In particular, the pacing degree between
spikes is determined in a statistical-mechanical way by quantifying the average contribution
of (microscopic) individual spikes to the (macroscopic) global potential VG. This "statistical-
mechanical" measure Ms is in contrast to the conventional measures such as the "thermo-
dynamic" order parameter (Golomb & Rinzel 1994; Hansel & Mato 2003), the "micro-
scopic" correlation-based measure (based on the cross-correlations between the microscopic
individual potentials) (Wang & Buzs´aki 1996; White et al. 1998), and the measures of pre-
cise spike timing based on the peri-stimulus time histogram (PSTH) (Mainen & Sejnowski
1995; Schreiber et al. 2003). The "thermodynamic" order parameter and the "microscopic"
measure concern just the the macroscopic global potential VG and the microscopic individual
potentials, respectively without considering any quantitative relation between VG and the mi-
croscopic individual potentials. (The auto-correlation of the global activity used in the work
of Brunel and Hakim (1999) may also be regarded as a kind of "thermodynamic" measure.)
For the PSTH-based measure "events," corresponding to peaks of the instantaneous popula-
tion firing rate, are selected through setting a threshold. Then, the measures for the reliability
and the precision of spike timing concern only the spikes within the events, in contrast to the
case of the "statistical-mechanical" measure where all spikes are considered (without select-
ing events). A main difference between the conventional and the new spike-based measures
lies in determining the pacing degree of spikes. The precision of spike timing for the con-
ventional case is given by just the standard deviation of (microscopic) individual spike times
within an event without considering the quantitative contribution of (microscopic) individ-
ual spikes to the (macroscopic) global activity. Hence, the PSTH-based measure is not a
4
statistical-mechanical measure. However, if we take the instantaneous population firing rate
as a global activity and exactly define the global cycles [see Fig. 4(a)] and the global phases
[see Eqs. (15) and (16)] like our case, then the conventional PSTH-based measure may
also develop into a similar statistical-mechanical measure. By varying the noise intensity,
we quantitatively characterize the stochastic spiking coherence in terms of the "statistical-
mechanical" measure Ms, and find that Ms reflects the degree of collective spiking coherence
seen in the raster plot very well. We also expect that Ms may be implemented for character-
izing the degree of collective coherence in the experimentally-obtained raster plot of neural
spikes. Finally, a summary is given in Section 4.
2 Morris-Lecar Neuron Model
In this section, we describe the biological neuron model used in our computational study.
We consider an ensemble of N globally coupled neurons. As an element in our neural sys-
tem, we choose the conductance-based ML neuron model with voltage-gated ion channels,
originally proposed to describe the time-evolution pattern of the membrane potential for
the giant muscle fibers of barnacles (Morris & Lecar 1981; Rinzel & Ermentrout 1998;
Tsumoto et al. 2006). The population dynamics in this neural network is governed by a set
of the following differential equations:
C
dvi
dt
dwi
dt
dsi
dt
= −Iion,i + IDC + Dx i − Isyn,i,
= f (w¥ (vi) − wi)
,
t R(vi)
= a s¥ (vi)(1 − si) − b si,
i = 1, · · · , N,
where
Iion,i = ICa,i + IK,i + IL,i
= gCam¥ (vi)(vi −VCa) + gKwi(vi −VK ) + gL(vi −VL),
Isyn,i =
J
N − 1
N(cid:229)
j(6=i)
s j(t)(vi −Vsyn),
m¥ (v) = 0.5 [1 + tanh {(v −V1)/V2}] ,
w¥ (v) = 0.5 [1 + tanh {(v −V3)/V4}] ,
t R(v) = 1/cosh {(v −V3)/(2V4)} ,
s¥ (vi) = 1/[1 + e−(vi−v∗)/d
].
(1)
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
Here, the state of the ith neuron at a time t (measured in units of ms) is characterized by
three state variables: the membrane potential vi (measured in units of mV), the slow recov-
ery variable wi representing the activation of the K+ current (i.e., the fraction of open K+
channels), and the synaptic gate variable si denoting the fraction of open synaptic ion chan-
nels. In Eq. (1), C represents the capacitance of the membrane of each neuron, and the time
evolution of vi is governed by four kinds of source currents.
The total ionic current Iion,i of the ith neuron consists of the calcium current ICa,i, the
potassium current IK,i, and the leakage current IL,i. Each ionic current obeys Ohm's law. The
constants gCa, gK, and gL are the maximum conductances for the ion and leakage channels,
5
and the constants VCa, VK, and VL are the reversal potentials at which each current is bal-
anced by the ionic concentration difference across the membrane. Since the calcium current
ICa,i changes much faster than the potassium current IK,i, the gate variable mi for the Ca2+
channel is assumed to always take its saturation value m¥ (vi). On the other hand, the activa-
tion variable wi for the K+ channel approaches its saturation value w¥ (vi) with a relaxation
time t R(vi)/f , where t R has a dimension of ms and f
is a (dimensionless) temperature-like
time scale factor.
Each ML neuron is also stimulated by the common DC current IDC and an indepen-
dent Gaussian white noise x [see the 2nd and 3rd terms in Eq. (1)] satisfying hx i(t)i = 0 and
hx i(t) x j(t′)i = d i j d (t −t′), where h· · ·i denotes the ensemble average. The noise x
is a para-
metric one which randomly perturbs the strength of the applied current IDC, and its intensity
is controlled by the parameter D. The last term in Eq. (1) represents the coupling of the
network. Each neuron is connected to all the other ones through global synaptic couplings.
Isyn,i of Eq. (6) represents such synaptic current injected into the ith neuron. Here the cou-
pling strength is controlled by the parameter J and Vsyn is the synaptic reversal potential. We
use Vsyn = −80 mV for the inhibitory synapse and Vsyn = 0 mV for the excitatory synapse.
The synaptic gate variable s obeys the 1st order kinetics of Eq. (3) (Golomb & Rinzel 1994;
Wang & Buzs´aki 1996). Here, the normalized concentration of synaptic transmitters, acti-
vating the synapse, is assumed to be an instantaneous sigmoidal function of the membrane
potential with a threshold v∗ in Eq. (10), where we set v∗ = 0 mV and d = 2 mV. The
transmitter release occurs only when the neuron emits a spike (i.e., its potential v is larger
than v∗). The synaptic channel opening rate, corresponding to the inverse of the synaptic rise
time t r, is a = 10 ms−1, which assures a fast rise of Isyn (Borgers & Kopell 2003, 2005). On
the other hand, the synaptic closing rate b , which is the inverse of the synaptic decay time
t d, depends on the type of the synaptic receptors. For the inhibitory GABAergic synapse
(involving the GABAA receptors) we set b = 0.1 ms−1, and for the excitatory glutamate
synapse (involving the AMPA receptors), we use b = 0.5 ms−1 (Borgers & Kopell 2003,
2005). Thus, Isyn decays slowly.
The ML neuron may exhibit either type-I or type-II excitability, depending on the sys-
tem parameters (Rinzel & Ermentrout 1998). For the case of type-I (type-II), the firing fre-
quency begins to increase from zero (a non-zero finite value) when IDC passes a threshold.
Throughout this paper, we consider the case of type-II excitability where gCa = 4.4 mS/cm2, gK =
8 mS/cm2, gL = 2 mS/cm2, VCa = 120 mV, VK = −84 mV, VL = −60 mV, C = 20 m F/cm2,
f = 0.04, V1 = −1.2 mV, V2 = 18 mV, V3 = 2 mV, and V4 = 30 mV. As IDC passes a thresh-
old in the absence of noise, each single type-II ML neuron begins to fire with a nonzero fre-
quency that is relatively insensitive to the change in IDC (Hodgkin 1948; Izhikevich 2000).
Numerical integration of Eqs. (1-3) is done using the Heun method (San Miguel & Toral
2000) (with the time step D t = 0.01 ms) similar to the second-order Runge-Kutta method,
and data for (vi, wi, si) (i = 1, . . . , N) are obtained with the sampling time interval D t = 1 ms.
For each realization of the stochastic process in Eqs. (1-3), we choose a random initial point
[vi(0), wi(0), si(0)] for the ith (i = 1, . . . , N) neuron with uniform probability in the range of
vi(0) ∈ (−70, 50), wi(0) ∈ (0.0, 0.6), and si(0) ∈ (0.0, 1.0).
6
)
V
m
(
v
50 (a)
-5
-60
1000
0.16 (b)
y
t
i
l
i
b
a
b
o
r
P
0.08
0.00
1500
t (ms)
2000
0
250
ISI (ms)
500
)
V
m
(
v
50 (c)
-5
-60
1000
1500
t (ms)
2000
Fig. 1 Intermittent noise-induced firings of a single subthreshold ML neuron: (a) time series of the membrane
potential v and (b) interspike interval (ISI) histogram for IDC = 87 m A/cm2 and D = 20 m A · ms1/2/cm2; the
ISI histogram is made of 5 × 104 ISIs and the bin size for the histogram is 5 ms. Regular firings of a single
suprathreshold ML neuron: (c) time series of v for IDC = 95 m A/cm2 and D = 0.
3 Characterization of Stochastic Spiking Coherence in An Inhibitory Population of
Subthreshold ML Neurons
In this section, we study stochastic spiking coherence in a population of inhibitory sub-
threshold ML neurons. The stochastic spiking coherence is quantitatively characterized in
terms of a new "statistical-mechanical" spike-based measure.
We first consider the case of a single ML neuron. An isolated subthreshold neuron cannot
fire spontaneously without noise; it may generate firings only with the aid of noise. Figure
1(a) shows a time series of the membrane potential v of a subthreshold neuron for IDC = 87
m A/cm2 and D = 20 m A · ms1/2/cm2. Noise-induced spikings appear intermittently. For
this subthreshold case, the ISI histogram is shown in Fig. 1(b). The most probable value of
the ISIs (corresponding to the main highest peak) is 97.5 ms. But, due to a long tail in the
ISI histogram the average value of the ISIs becomes 161.6 ms. This noise-induced irregular
oscillation of v is in contrast to the regular oscillation of v for a suprathreshold neuron (which
fires spontaneously without noise) when IDC = 95 m A/cm2 and D = 0 [see Fig. 1(c)].
We consider an inhibitory population of N globally coupled subthreshold ML neurons
for IDC = 87 m A/cm2 and set the coupling strength as J = 3 mS/cm2. (Hereafter, for con-
venience we omit the dimensions of IDC, D, and J.) By varying the noise intensity D, we
investigate the stochastic spiking coherence. Emergence of collective spiking coherence may
be well described by the (population-averaged) global potential,
VG(t) =
1
N
N(cid:229)
i=1
vi(t).
(11)
In the thermodynamic limit (N → ¥ ), a collective state becomes coherent if D VG(t) (=
VG(t) − VG(t)) is non-stationary (i.e., an oscillating global potential VG appears for a co-
herent case), where the overbar represents the time average. Otherwise (i.e., when D VG is
stationary), it becomes incoherent. Thus, the mean square deviation of the global potential
7
2.0
(a)
-1.5
O
0
1
g
o
l
-5.0
0.5
(b)
1000
i
500
1500
t (ms)
0
10
0
1000
(c)
50
-5
s
N
)
V
m
(
G
V
1.0
1.5
log10D
(d)
1000
500
i
0
100
s
N
2000
)
V
m
(
G
V
0
1000
(e)
50
-5
N=102
N=103
N=104
2.0
1500
t (ms)
2000
-60
1000
1500
t (ms)
2000
-60
1000
1500
t (ms)
2000
Fig. 2 Order parameter and partial occupation in N(= 103) globally coupled inhibitory subthreshold ML
neurons: (a) plots of the order parameter O versus the noise intensity D. (b) partially-occupied raster plot
of spikings a (i: neuron index) and plot of number of spikes (Ns) versus time (t) and (c) time series of the
ensemble-averaged global potential VG. Full occupation in N(= 103) globally coupled excitatory ML neurons:
(d) fully-occupied raster plot of spikings and plot of Ns versus t and (e) time series of VG. For both inhibitory
and excitatory cases, IDC = 87 m A/cm2, D = 20 m A · ms1/2/cm2, J = 3 mS/cm2, and the bin size for Ns in
(b) and (d) is 1 ms
VG (i.e., time-averaged fluctuations of VG),
O ≡ (VG(t) −VG(t))2,
(12)
plays the role of an order parameter used for describing the coherence-incoherence tran-
sition (Manrubia et al. 2004). For the coherent (incoherent) state, the order parameter O
approaches a non-zero (zero) limit value as N goes to the infinity. Figure 2(a) shows a plot
of the order parameter versus the noise intensity. For D < D∗
l (≃ 9.4), incoherent states exist
because the order parameter O tends to zero as N → ¥
. As D passes the lower threshold D∗
l ,
a coherent transition occurs because of a constructive role of noise to stimulate coherence
between noise-induced spikings. However, for large D > D∗
h (≃ 33.4) such coherent states
disappear (i.e., a transition to an incoherent state occurs when D passes the higher threshold
D∗
h) due to a destructive role of noise to spoil the collective spiking coherence.
8
As an example, we consider a coherent state for D = 20. For this case stochastic spiking
coherence may be well visualized in terms of the raster plot of neural spikes. The upper
panel in Fig. 2(b) shows the raster plot for N = 103. We note that stripes (composed of spikes
and indicating collective coherence) appear successively in the raster plot. The number of
spikes (Ns) in each stripe is given in the lower panel. About 100 spikes constitute a stripe
[i.e., only a fraction (about 1/10) of the total neurons fire in each stripe]. In this way partial
occupation occurs in the stripes. A regularly oscillating global potential VG emerges through
cooperation of spikes in partially-occupied stripes. A time series of VG is shown in Fig. 2(c).
The amplitude of VG is much smaller than that in Fig. 1(a) for the isolated single case.
This reduction of amplitude occurs mainly because of partial occupation. Local maxima
of VG appear at the centers of stripes where the spiking densities are locally highest. The
global period TG of VG (corresponding to the average intermax interval of VG or equivalently
corresponding to the average interstripe interval in the raster plot) is 54.2 ms. Hence the
global frequency fG (= 18.5 Hz) of VG is roughly as twice as the most probable frequency
(= 10.3 Hz) of v for the isolated single case. Thus, a regular global oscillation with reduced
amplitude and increased frequency occurs for the partially-occupied case. For comparison,
we also consider the case of full occupation which occurs in a population of excitatory
neurons for the same set of parameters IDC, D, and J. [We emphasize that partial occupation
may also occur even for the excitatory case of smaller J (e.g., J = 1).] Figures 2(d) and
2(e) show the raster plot and a time series of VG for the fully-occupied case, respectively.
Fully-occupied stripes appear successively at nearly regular time intervals D t (= 97.9 ms),
in contrast to the partially-occupied case. A regularly oscillating VG with large amplitude
appears via cooperation of spikes in the fully-occupied stripes, and its global frequency fG
(= 10.2 Hz) is smaller than that for the partially-occupied case (in fact, fG for the fully-
occupied case is nearly equal to the most probable frequency of v for the isolated single
case).
To further understand the full and the partial occupations in the raster plots, we examine
the local and the global output signals for both the fully-occupied and the partially-occupied
cases of D = 20. Since our neural network is globally coupled, any local neuron may be
a representative one. Unlike the isolated single case (shown in Fig. 1), each local neuron
within the network is stimulated by a synaptic current which is directly related to the aver-
age of firing events of all the other neurons. First, we consider the fully-occupied excitatory
case, and investigate the time series of the local potential v1 of the 1st neuron and the global
potential VG which are shown in Fig. 3(a). The 1st neuron fires spikes phase-locked to VG at
every cycle of VG (i.e., the local potential v1 exhibits a nearly regular phase-locked oscilla-
tion). To confirm this 1:1 phase locking, we collect 5 × 104 ISIs from all local neurons, and
obtain the ISI histogram [see Fig. 3(b)]. A single peak with a small width is located around
the average value (= 97.9 ms) which agrees well with the global period TG of VG. This 1:1
phase locking is expected to occur for the fully-occupied case because the global frequency
fG (= 10.2 Hz) of VG is nearly equal to the most probable frequency (= 10.3 Hz) of noise-
induced oscillation for the uncoupled single neuron. As a result of such 1:1 phase locking,
full occupation occurs in the raster plot. Second, we study the partially-occupied inhibitory
case. Figure 3(c) shows the time series of the local potential v1 of the 1st neuron and the
global potential VG. The 1st neuron exhibits intermittent spikings phase-locked to VG at ran-
dom multiples of the period TG (=54.2 ms) of VG. In addition to these intermittent spiking
phases, hopping phases (exhibiting small subthreshold oscillations) appear in most of global
cycles. We also note that after occurrence of a spiking, recovery from a hyperpolarized to
a resting state is made during the next global cycle. Hence, a "preparatory" phase without
spiking and hopping (for preparing for generation of the next spike or hopping) follows
1500
t (ms)
(c)
30
-40
30
-40
)
V
m
(
1
v
)
V
m
(
G
V
2000
1000
y
t
i
l
i
b
a
b
o
r
P
0.08 (d)
0.04
0.00
0
9
2000
1500
t (ms)
500
ISI (ms)
1000
(a)
30
)
V
m
(
1
v
-40
30
)
V
m
(
G
V
-40
1000
0.4 (b)
y
t
i
l
i
b
a
b
o
r
P
0.2
0.0
0
100
ISI (ms)
200
Fig. 3 1:1 phase locking (spikings phase-locked to the global potential VG at every cycle of VG) in N(=
103) globally coupled excitatory ML neurons: (a) time series of the local potential v1 of the 1st neuron and
the ensemble-averaged global potential VG and (b) interspike interval (ISI) histogram with a single peak.
Stochastic phase locking leading to stochastic spike skipping (intermittent spikings phase-locked to VG at
random multiples of the period of VG) in N(= 103) globally coupled inhibitory subthreshold ML neurons: (c)
time series of v1 and VG, and (d) ISI histogram with multiple peaks. For both inhibitory and excitatory cases,
IDC = 87 m A/cm2, D = 20 m A · ms1/2/cm2, and J = 3 mS/cm2. Vertical dashed lines in (a) and (c) represent
the times at which local minima of VG occur. Each ISI histogram in (b) and (d) is made of 5 × 104 ISIs and
the bin size for the histogram is 5 ms. Vertical dotted lines in (d) denote integer multiples of the global period
TG (=54.2 ms) of VG.
each spiking phase (see the gray parts). In this way, the local potential exhibits an irregular
firing pattern consisting of randomly phase-locked spiking and hopping phases and prepara-
tory phases. We note that a regular fast global oscillation with a small amplitude emerges
from these irregular individual oscillations. To confirm the stochastic spike skipping (arising
fromn stochastic phase locking) in the local potential, we collect 5 × 104 ISIs from all local
neurons and get the ISI histogram. Multiple peaks appear at multiples of the period TG of
the global potential VG. However, due to appearance of preparatory cycles, the 1st peak of
the histogram (which is the highest one) appears at 2TG (not TG). Hence, local neurons fire
mostly in alternate global cycles. Due to this stochastic spike skipping partial occupation
occurs in the raster plot. Similar skipping phenomena of spikings (characterized with multi-
peaked ISI histograms) were found in single noisy neuron models driven by a weak periodic
external force for the stochastic resonance (Longtin 1995, 2000; Gammaitoni et al. 1998).
Unlike this single case, stochastic spike skipping in networks of inhibitory subthreshold
neurons is a collective effect because it occurs due to a driving by a coherent ensemble-
averaged synaptic current. As discussed in details in Section 4, similar results on skipping
were also obtained in simplified networks of IF neurons (Brunel and Hakim 1999; Brunel
G1
G2
(b)
14
7
0
-
1040
1110
t (ms)
1180
G1
0
G2
2
3
1110
t (ms)
1180
i
(d)
i
O
i
P
i
M
1
0
1
0
1
0
2000
3000
0
1000
i
2000
3000
10
(a)
-20
)
V
m
(
-28
G
V
-36
1040
(c)
0.2
i
O
0.0
1
i
P
0
0.2
i
M
0.0
0
1000
Fig. 4 "Statistical-mechanical" spike-based coherence measure in N(= 103) globally coupled ML neurons
for IDC = 87 m A/cm2, D = 20 m A · ms1/2/cm2, and J = 3 mS/cm2. Partially-occupied inhibitory case: (a)
time series of the global potential VG, (b) plot of the global phase (F ) versus time (t), and (c) plots of Oi
(occupation degree of spikes in the ith stripe), Pi (pacing degree of spikes in the ith stripe), and Mi (spiking
measure in the ith stripe) versus i (stripe). In (a) and (b), vertical dashed and solid lines represent the times
at which local minima and maxima (denoted by open and solid circles) of VG occur, respectively and Gi
(i = 1,2) denotes the ith global cycle. Fully-occupied excitatory case: (d) plots of Oi, Pi, and Mi versus i.
2000; Brunel & Wang 2003). However, there are some differences in the spiking pattern of
individual neurons and the effect of noise on the collective coherence.
Our main purpose is to quantitatively characterize the stochastic spiking coherence
which is well visualized in the raster plot of spikes. To measure the degree of stochas-
tic spiking coherence seen in the raster plots in Fig. 2 for D = 20, we introduce a new
"statistical-mechanical" spike-based coherence measure Ms by considering the occupation
pattern and the pacing pattern of spikes in the stripes of the raster plot. Particularly, the
pacing degree between spikes is determined in a statistical-mechanical way by quantifying
the average contribution of microscopic individual spikes to the macroscopic global poten-
tial VG. The spiking coherence measure Mi of the ith stripe is defined by the product of the
occupation degree Oi of spikes (representing the density of the ith stripe) and the pacing
degree Pi of spikes (denoting the smearing of the ith stripe):
The occupation degree Oi in the ith stripe is given by the fraction of spiking neurons:
Mi = Oi · Pi.
Oi =
N(s)
i
N
,
(13)
(14)
11
i
where N(s)
is the number of spiking neurons in the ith stripe. For the full occupation, Oi = 1,
while for the partial occupation Oi < 1. The pacing degree Pi of each microscopic spike in
the ith stripe can be determined in a statistical-mechanical way by taking into consideration
its contribution to the macroscopic global potential VG. Figure 4(a) shows a time series of
the global potential VG; local maxima and minima are denoted by solid and open circles, re-
spectively. Central maxima of VG between neighboring left and right minima of VG coincide
with centers of stripes in the raster plot. The global cycle starting from the left minimum of
VG which appears first after the transient time (= 103 ms) is regarded as the 1st one, which
is denoted by G1. The 2nd global cycle G2 begins from the next following right minimum
of G1, and so on. Then, we introduce an instantaneous global phase F (t) of VG via linear
interpolation in the two successive subregions forming a global cycle (Freund et al. 2003).
The global phase F (t) between the left minimum (corresponding to the beginning point of
the ith global cycle) and the central maximum is given by:
F (t) = 2p (i − 3/2) + p t − t(min)
i
− t(min)
t(max)
i
i
! for t(min)
i
≤ t < t(max)
i
(i = 1, 2, 3, . . .),
(15)
and F (t) between the central maximum and the right minimum (corresponding to the be-
ginning point of the (i + 1)th cycle) is given by:
F (t) = 2p (i − 1) + p t − t(max)
i+1 − t(max)
t(min)
i
i
! for t(max)
i
≤ t < t(min)
i+1
(i = 1, 2, 3, . . .),
(16)
i
i
where t(min)
is the beginning time of the ith global cycle (i.e., the time at which the left
minimum of VG appears in the ith global cycle) and t(max)
is the time at which the maximum
of VG appears in the ith global cycle. The global phase F
in the first two global cycles
is shown in Fig. 4(b). Then, the contribution of the kth microscopic spike in the ith stripe
to VG is given by cosF k, where F k is the global phase at the
occurring at the time t(s)
k
kth spiking time [i.e., F k ≡ F (t(s)
k )]. A microscopic spike makes the most constructive (in-
phase) contribution to VG when the corresponding global phase F k is 2p n (n = 0, 1, 2, . . .),
while it makes the most destructive (anti-phase) contribution to VG when F
i is 2p (n − 1/2).
By averaging the contributions of all microscopic spikes in the ith stripe to VG, we obtain
the pacing degree of spikes in the ith stripe,
Pi =
1
Si
Si
k=1
cosF k,
(17)
where Si is the total number of microscopic spikes in the ith stripe. By averaging Mi of
Eq. (13) over a sufficiently large number Ns of stripes, we obtain the 'statistical-mechanical"
spike-based coherence measure Ms:
Ms =
1
Ns
Ns
i=1
Mi.
(18)
We follow 3 × 103 stripes and get Oi, Pi, and Mi in each ith stripe for the partially-
occupied inhibitory case of Fig. 2(b). The results are shown in Fig. 4(c). For comparison,
we also measure the degree of stochastic spiking coherence seen in the raster plot of Fig. 2(d)
for the fully-occupied excitatory case and give the results in Fig. 4(d). We note a distinct dif-
ference in the average occupation degree hOii, where h· · ·i denotes the average over stripes.
(cid:229)
(cid:229)
12
(a1)
1000
i
500
0
10
0
s
N
y
t
i
l
i
b
a
b
o
r
P
0.08
0.04
0.00
1200
(b1)
200
(c)
0.14
(a2)
1800
1200
(b2)
(a3)
1800 1200
t (ms)
(b3)
(a4)
1800 1200
(b4)
800
200
1.0
800
200
ISI (ms)
(d)
800
200
(e)
0.12
1800
800
i
>
O
<
0.07
i
>
P
<
0.5
s
M
0.06
0.00
0.90
1.25
log10D
0.0
0.90
1.60
1.25
log10D
0.00
0.90
1.60
1.60
1.25
log10D
Fig. 5 Stochastic spiking coherence for various values of D in N(= 103) globally coupled inhibitory sub-
threshold ML neurons when IDC = 87 m A/cm2 and J = 3 mS/cm2. (a) Raster plots of spikings (i: neuron in-
dex) and plots of number of spikes (Ns) versus time (t) and (b) interspike interval (ISI) histograms for (a1) and
(b1) D = 10 m A · ms1/2/cm2, (a2) and (b2) D = 20 m A · ms1/2/cm2, (a3) and (b3) D = 30 m A · ms1/2/cm2,
and (a4) and (b4) D = 32 m A · ms1/2/cm2. The bin size for Ns in (a) is 1 ms. Each ISI histogram in (b) is
made of 5 × 104 ISIs and the bin size for the histogram is 5 ms. Vertical dotted lines in (b) represent the
integer multiples of the global period TG of VG; TG = (b1) 67.4 ms, (b2) 54.2 ms, (b3) 48.6 ms, and (b4) 47.7
ms. (c) Plot of hOii (average occupation degree of spikes) versus log10 D. (d) Plot of hPii (average pacing
degree of spikes) versus log10 D. (e) Plot of Ms ("statistical-mechanical" spike-based coherence measure)
versus log10 D. To obtain hOii, hPii, and Ms in (c)-(e), we follow the 3 × 103 stripes for each D. Open circles
in (c)-(e) denote the data for D = 20 m A · ms1/2/cm2.
For the inhibitory case, partial occupation with very small hOii (= 0.106) occurs due to
stochastic spike skipping, while for the excitatory case full occupation with hOii = 1 occurs
as a result of 1:1 phase locking. For the partially-occupied case of inhibitory coupling, the
average pacing degree hPii (= 0.766) is large in contrast to hOii, although it is smaller than
hPii (= 0.911) for the fully-occupied case of excitatory coupling. As a result, the "statistical-
mechanical" coherence measure Ms (which represents the collective coherence seen in the
whole raster plot) is 0.081 for the partially-occupied inhibitory case which is very low when
compared with Ms (= 0.911) for the fully-occupied excitatory case. The main reason for the
low degree of stochastic spiking coherence is mainly due to partial occupation.
Finally, by varying D we characterize the stochastic spiking coherence in terms of the
"statistical-mechanical" measure Ms in a population of inhibitory subthreshold ML neurons
when IDC = 87 and J = 3. Such stochastic spiking coherence may be well visualized in the
raster plots of neural spikes. Figures 5(a1)-5(a4) show the raster plots and the temporal plots
of the number of spikes (Ns) for N = 103 in the coherent region for D = 10, 20, 30, and
13
32, respectively. The corresponding ISI histograms are also given in Figs. 5(b1)-5(b4). We
measure the degree of stochastic spiking coherence in terms of hOii (average occupation
degree), hPii (average pacing degree), and Ms for 20 values of D in the coherent regime, and
the results are shown in Figs. 5(c)-5(e). (Here, for comparison with other values of D we
include the case of D = 20 which is studied in details above; open circles in Figs. 5(c)-5(e)
represent the data for D = 20.) For the coherent case of D=20, the raster plot in Fig. 5(a2)
consists of relatively clear partially-occupied stripes with hOii = 0.106 and hPii = 0.766.
Such partial occupation results from stochastic spike skipping of individual neurons seen
well in the ISI histogram of Fig. 5(b2) with clear (well-separated) multiple peaks appear-
ing at nTG (TG: global period of VG and n=2,3,...). As the value of D is increased from 20,
the average occupation degree hOii increases only a little, as might be seen from the raster
plots and the plots of Ns in Figs. 5(a3) and 5(a4) (hOii = 0.114 and 0.115 for D = 30 and
32, respectively). This slow increase in hOii is well shown in Fig. 5(c). However,the aver-
age pacing degree hPii for D > 20 decreases rapidly, as shown in Fig. 5(d). For example,
stripes in the raster plots of Figs. 5(a3) and 5(a4) become more and more smeared, and
hence the average pacing degree is decreased with increase in D. This smearing of stripes
can be understood from the change in the structure of the ISI histograms. As D is increased,
the heights of peaks are decreased, but their widths are widened [see Figs. 5(b3) and 5(b4)].
Thus, peaks begin to merge. This merging of peaks results in the smearing of stripes. We
note that for large D merged multiple peaks in the ISI histogram are directly associated with
smeared partially-occupied stripes in the raster plot. Thus, for D > 20 the degree of stochas-
tic spiking coherence is rapidly decreased as shown in Fig. 5(e), mainly due to the rapid
decrease in hPii (a little increase in hOii has negligible effect). Eventually, when passing the
higher threshold D∗
h (≃ 33.4) stripes no longer exist due to complete smearing, and then
incoherent states appear.
By decreasing the value of D from 20, we also characterize the stochastic spiking co-
herence. As an example see the raster plot of spikes in Fig. 5(a1) for D = 10. The average
occupation degree is much decreased to hOii = 0.044, as can also be expected from the plot
of Ns. This rapid decrease in hOii for D < 20 can be seen in Fig. 5(c). Contrary to hOii the
average pacing degree hPii (= 0.684) for D = 10 is decreased a little when compared to
the value of hPii (= 0.766) for D = 20; only a little more smearing occurs. For this case,
the ISI histogram has multiple peaks without merging like the case of D=20, as shown in
Fig. 5(b1). However, when compared to the case of D = 20 the heights of peaks decrease,
and the average ISI increases via appearance of long ISIs. Thus, the degree of stochastic
spiking coherence for D = 10 (Ms = 0.032) is much decreased mainly due to rapid decrease
in the average occupation degree hOii (small decrease in hPii has only a little effect). In fact,
as can be seen in Fig. 5(d) there is no noticeable change in hPii for 10 < D < 20; as D is
decreased from 20 to a value of D (≃ 14), hPii increases slowly, and then it begins to de-
crease. Thus, for D < 10 both hOii and hPii decreases so rapidly, as shown in Figs. 5(c) and
5(d). Hence, as D is decreased from 10 the degree of stochastic spiking coherence decreases
rapidly due to both effects of hOii and hPii. Eventually, when D is decreased through the
lower threshold D∗
l (≃ 9.4), completely scattered sparse spikes appear without forming any
stripes in the raster plot, and thus incoherent states exist for D < D∗
l . In this way, we char-
acterize the stochastic spiking coherence in terms of the "statistical-mechanical" measure
Ms in the whole coherent region, and find that Ms reflects the degree of collective coherence
seen in the raster plot very well. When taking into consideration both the occupation degree
and the pacing degree of spikes in the raster plot, a maximal spiking coherence occurs near
D ≃ 20 [see Fig. 5(e)]. As discussed in details in Section 1, Ms is in contrast to the conven-
tional measures where any quantitative relation between (microscopic) individual spikes and
14
the (macroscopic) global potential VG is not considered [e.g., the "thermodynamic" order
parameter (Golomb & Rinzel 1994; Hansel & Mato 2003), the "microscopic" correlation-
based measure (Wang & Buzs´aki 1996; White et al. 1998), and the PSTH-based measures
of precise spike timing (Mainen & Sejnowski 1995; Schreiber et al. 2003)].
4 Summary
By changing the noise intensity, we have characterized stochastic spiking coherence in an in-
hibitory population of subthreshold biological ML neurons. In a range of intermediate noise
intensity, a regularly oscillating global potential VG (with reduced amplitude and the in-
creased frequency) emerges via cooperation of irregular individual firing activities. We note
that this stochastic spiking coherence may be well visualized in the raster plot of spikes. For
the case of a coherent state, partially-occupied stripes appear in the raster plot. This partial
occupation occurs due to stochastic spike skipping, which is shown clearly in the multi-
peaked ISI histogram. Recently, Brunel and Hakim (1999) have obtained similar results on
spike skipping in a simplified network of integrate-and-fire (IF) neurons which are randomly
connected via instantaneous inhibitory synapses (modeled by the delta function with a trans-
mission delay). [Such skipping was also reported in the network of inhibitory and excitatory
IF neurons (Brunel 2000; Brunel & Wang 2003).] For the case of the IF neuron a spike is
triggered instantaneously when the membrane potential reaches a threshold, in contrast to
the case of the biological ML neuron where dynamics of voltage-gated ion channels leads to
spike generation. As the stimulus exceeds a threshold, the firing frequency of the IF neuron
begins to increase from zero (i.e., the IF neuron exhibits the type-I excitability), unlike the
case of the type-II ML neuron (used in our computational study). These type-I IF neurons are
driven by random excitatory inputs activated by independent Poisson processes. In this sim-
plified network, it was shown that weakly coherent global activities (roughly corresponding
to the global potential VG) emerge from irregular firing patterns of neurons [refer to Figure 3
in the paper of Brunel and Hakim (1999)]. The collective coherence was well shown in the
temporal auto-correlation function of the global activity. Although the basic results in the
simplified network are similar to ours in a realistic network, there are some differences in the
spiking pattern of individual neurons. The stochastic spike emission of individual IF neurons
was well shown in the Poissonian ISI histogram showing little indication of the collective be-
havior. Since no clear multiple peaks appear in the Poisson-like ISI histogram, the individual
IF neurons exhibit more stochastic spike emission than the ML neurons, although weak col-
lective coherence occurs in both cases. There are some additional differences in the effect of
the external noise on the collective coherence. As the magnitude of the external noise is in-
creased from zero and passes a threshold, the global activity was found to exhibit a transition
from an oscillatory to a stationary state (Brunel and Hakim 1999). Thus, incoherent states
appear for the case of strong noise. On the other hand, through competition between the
constructive role (stimulating coherence between noise-induced spikes) and the destructive
role (spoiling the collective coherence) of noise, stochastic spiking coherence (in our work)
appears in a range of intermediate noise intensities. Hence, incoherent states appear for both
cases of sufficiently strong and weak noise. These differences in both the spiking pattern
of individual neurons and the noise effect on the collective coherence seem to arise mainly
from different types of drivings on individual neurons; IF neurons are driven by random ex-
citatory inputs activated by Poisson processes, while ML neurons are driven by a Gaussian
white noise in addition to a subthreshold DC stimulus. The main purpose of our work is
to quantitatively measure the degree of stochastic spiking coherence seen in the raster plot.
15
We have introduced a new type of spike-based coherence measure Ms by taking into con-
sideration the occupation degree and the pacing degree of spikes in the stripes. Particularly,
the pacing degree between spikes is determined in a statistical-mechanical way by quantify-
ing the average contribution of (microscopic) individual spikes to the (macroscopic) global
potential VG. This "statistical-mechanical" measure Ms is in contrast to the conventional
measures (e.g., the "thermodynamic" order parameter, the "microscopic" correlation-based
measure, and the PSTH-based measures of precise spike timing) where any quantitative re-
lation between the microscopic individual spikes and the macroscopic global potential VG
is not considered. In terms of Ms, we have quantitatively characterized the stochastic spik-
ing coherence, and found that Ms reflects the degree of collective spiking coherence seen in
the raster plot very well. Finally, we also expect that Ms may be implemented to character-
ize the degree of collective spiking coherence in an experimentally-obtained raster plot of
neural spikes.
Acknowledgements This research was supported by the Basic Science Research Program through the Na-
tional Research Foundation of Korea funded by the Ministry of Education, Science and Technology (2009-
0070865). S.-Y. Kim thanks Dr. D.-G. Hong for his interest in our work.
References
Borgers, C. & Kopell, N. (2003). Synchronization in networks of excitatory and inhibitory neurons with
sparse, random connectivity. Neural Computation, 15, 509-538.
Borgers, C. & Kopell, N. (2005). Effects of noisy drive on rhythms in networks of excitatory and inhibitory
neurons. Neural Computation, 17, 557-608.
Brunel, N. & Hakim, V. (1999). Fast global oscillations in networks of intergrate-and-fire neurons with low
firing rates. Neural Computation, 11, 1621-1671.
Brunel, N. (2000). Dynamics of sparsely connected networks of excitatory and inhibitory spiking neurons.
Journal of Computational Neuroscience, 8, 183-208.
Brunel, N. & Wang, X.-J. (2003). What determines the frequency of fast network oscillations with irregular
neural discharge? I. Snaptic dynamics and excitation-inhibition balance. Journal of Neurophysiology, 90,
415-430.
Buzs´aki, G., Geisler, C., Henze, D. A., & Wang, X.-J. (2004). Interneuron diversity series: Circuit complexity
and axon wiring economy of cortical interneurons. Trends in Neuroscience, 27, 186-193.
Buzs´aki, G. (2006). Rhythms of the Brain. New York: Oxford University Press.
Freund, J., Schimansky-Geier, L., & Hanggi, P. (2003). Frequency and phase synchronization in stochastic
systems. Chaos, 13, 225-238.
Gammaitoni, L., Hanggi, P., Jung, P., & Marchesoni, F. (1998). Stochastic resonance. Reviews of Modern
Physics, 70, 223-287.
Golomb, D. & Rinzel, J. (1994). Clustering in globally coupled inhibitory neurons. Physica D, 72, 259-282.
Gray, C. M. (1994). Synchronous oscillations in neuronal systems: Mechanisms and functions. Journal of
Computational Neuroscience, 1, 11-38.
Grosse, P., Cassidy, M. J., & Brown, P. (2002). EEG-EMG, MEG-EMG and EMG-EMG frequency analysis:
Physiological principles and clinical applications. Clinical Neurophysiology, 113, 1523-1531.
Hansel, D., Mato, G., & Meunier, C. (1995). Synchrony in excitatory neural networks. Neural Computation,
7, 307-337.
Hansel, D. & Mato, G. (2003). Asynchronous states and the emergence of synchrony in large networks of
interacting excitatory and inhibitory neurons. Neural Computation, 15, 1-56.
Hauschildt, B., Janson, N. B., Balanov, A., & Scholl, E. (2006). Noise-induced cooperative dynamics and its
control in coupled neuron models. Physical Review E, 74, 051906.
Hodgkin, A. L. (1948). The local electric changes associated with repetitive action in a non-medullated axon.
Journal of Physiology, 107, 165-181.
Izhikevich, E. M. (2000). Neural excitability, spiking and bursting. International Journal of Bifurcation and
Chaos, 10, 1171-1266.
Kuramoto, Y. (1991). Collective synchronization of pulse-coupled oscillators and excitable units. Physica D,
50, 15-30.
16
Lim, W. & Kim, S.-Y. (2007). Characterization of stochastic spiking coherence in coupled neurons. Journal
of the Korean Physical Society, 51, 1427-1431.
Lim, W. & Kim, S.-Y. (2009). Stochastic spiking coherence in coupled subthreshold Morris-Lecar neurons.
International Journal of Modern Physics B, 23, 703-710.
Longtin, A. (1995). Synchronization of the stochastic Fitzhugh-Nagumo equations to periodic forcing. Nuovo
Cimento D, 17, 835-846.
Longtin, A. (2000). Stochastic aspects of neural phase locking to periodic signals. In Kim S., Lee K. J., &
Sung W. (Eds.), Stochastic Dynamics and Pattern Formation in Biological and Complex Systems (pp.
219-239). New York: AIP.
Mainen, Z. F. & Sejnowski T. J. (1995). Reliability of spike timing in neocortical neurons. Science, 268,
1503-1506.
Manrubia, S. C., Mikhailov, A. S., & Zanette, D. H. (2004). Emergence of Dynamical Order. Singapore:
World Scientific.
Mirollo, R. E. & Strogatz, S. H. (1990). Synchronization of pulse-coupled biological oscillators. SIAM Jour-
nal on Applied Mathematics, 50, 1645-1662.
Morris, C. & Lecar, H. (1981). Voltage oscillations in the barnacle giant muscle fiber. Biophysical Journal,
35, 193-213.
Rinzel, J. & Ermentrout, B. (1998). Analysis of neural excitability and oscillations. In Koch, C. & Segev, I.
(Eds.), Methods in Neural Modeling: from Ions to Networks (pp. 251-292). Cambridge: MIT Press.
San Miguel, M. & Toral, R. (2000). Stochastic effect in physical systems. In Martinez, J., Tiemann, R., &
Tirapegui, E. (Eds.), Instabilities and Nonequilibrium Structures VI (pp. 35-130) Dordrecht: Kluwer
Academic Publisher.
Schreiber, S., Fellous, J. M., Whitmer, D., Tiesinga, P., & Sejnowski, T. J. (2003), A new correlation-based
measure of spike timing reliability. Neurocomputing, 52-54, 925-931.
Tass, P., Rosenblum, M. G., Weule, J., Kurths, J., Pikovsky, A., Volkmann, J., Schnitzler, A., & Freud, H.
J. (1998). Detection of n : m phase locking from noisy data: application to magnetoencephalography.
Physical Review Letters, 81, 3291-3294.
Tiesinga, P. H. E., Fellous, J.-M., Jose, J. V., & Sejnowski, T. J. (2001). Computational model of carbachol-
induced delta, theta, and gamma oscillations in the hippocampus. Hippocampus, 11, 251-274.
Tsumoto, K., Kitajima, H., Yoshinaga, T., Aihara, K., & Kawakami, H. (2006). Bifurcations in Morris-Lecar
neuron model. Neurocomputing, 69, 293-316.
Vreeswijk, C. van, Abbott, L. F., & Ermentrout, G. B. (1994). When inhibition not excitation synchronizes
neural firing. Journal of Computational Neuroscience, 1, 313-321.
Wang, X.-J. & Rinzel, J. (1992). Alternating and synchronous rhythms in reciprocally inhibitory model neu-
rons. Neural Computation, 4, 84-97.
Wang, X.-J. & Buzs´aki, G. (1996). Gamma oscillations by synaptic inhibition in a hippocampal interneuronal
network. The Journal of Neuroscience, 16, 6402-6413.
Wang, X.-J. (2003). Neural oscillations. In Nadel, L. (Ed.), Encyclopedia of Cognitive Science (pp. 272-280).
London: MacMillan.
Wang, Y., Chik, D. T. W., & Wang, Z. D. (2000). Coherence resonance and noise-induced synchronization in
globally coupled Hodgkin-Huxley neurons. Physical Review E, 61, 740-746.
White, J., Chow, C. C., Ritt, J., Soto-Trevino, C., & Kopell, N. (1998). Synchronization and oscillatory
dynamics in heterogeneous, mutually inhibited neurons. Journal of Computational Neuroscience, 5, 5-
16.
Whittington, M. A., Traub, R. D., Kopell, N., Ermentrout, B., & Buhl, E. H. (2000). Inhibition-based rhythms:
experimental and mathematical observations on network dynamics. International Journal of Psychophys-
iology, 38, 315-336.
|
1010.0617 | 1 | 1010 | 2010-10-04T15:06:08 | Bone in vivo: Surface mapping technique | [
"physics.bio-ph",
"physics.med-ph"
] | Bone surface mapping technique is proposed on the bases of two kinds of uniqueness of bone in vivo, (i) magnitude of the principal moments of inertia, (ii) the direction cosines of principal axes of inertia relative to inertia reference frame. We choose the principal axes of inertia as the bone coordinate system axes. The geographical marks such as the prime meridian of the bone in vivo are defined and methods such as tomographic reconstruction and boundary development are employed so that the surface of bone in vivo can be mapped. Experimental results show that the surface mapping technique can both reflect the shape and help study the surface changes of bone in vivo. The prospect of such research into the surface shape and changing laws of organ, tissue or cell will be promising. | physics.bio-ph | physics |
1
Bone in vivo: Surface mapping technique
Yifang Fan1,∗, Yubo Fan2, Zhiyu Li3, Changsheng Lv1
1 Center for Scientific Research, Guangzhou Institute of Physical Education, Guangzhou
510500, P.R. China
2 Bioengineering Department, Beijing University of Aeronautics and Astronautics, Beijing
100191, P.R. China
3 College of Foreign Languages, Jinan University, Guangzhou 510632, P.R. China
∗ E-mail: [email protected]
Abstract
Abstract: Bone surface mapping technique is proposed on the bases of two kinds of uniqueness of bone
in vivo, (i) magnitude of the principal moments of inertia, (ii) the direction cosines of principal axes of
inertia relative to inertia reference frame. We choose the principal axes of inertia as the bone coordinate
system axes. The geographical marks such as the prime meridian of the bone in vivo are defined and
methods such as tomographic reconstruction and boundary development are employed so that the surface
of bone in vivo can be mapped. Experimental results show that the surface mapping technique can both
reflect the shape and help study the surface changes of bone in vivo. The prospect of such research into
the surface shape and changing laws of organ, tissue or cell will be promising.
Introduction
The shape of bone is the adaptive result of bone in the mechanical and physiological environment [1–4].
A map, the visual representation of our real world symbol model, can reveal not only the spatial structure
properties of an object but also the changes in time series [5–7]. The mapping of bone surface, therefore,
is used as an approach to study the adaptability of bone morphology. Mapping technique has become
an even more powerful and useful method to do scientific research. Mapping and flattening techniques
have also been widely used in medical research [8–13]. They are both concerned with the development
methods. The flattening technique develops the three-dimensional object to a two-dimensional one [14]
while the mapping technique plays an essential role in interpreting the surface structure of an object [15].
The advancements and improvements of three-dimensional imaging of bone in vivo [16–18] have brought
better data collection methods of bone surface, but the mapping techniques have not been systematically
explored with satisfactory results.
To map the bone surface, some geographic marks and geographic coordinates such as the bone surface
prime meridian, the equator line or the contour should be identified [19]. They should be defined on the
In order to study the bone’s changes caused by the external factors, it is
bone’s coordinate system.
fundamental to set up a bone coordinate system when mapping the bone surface.
In our study, we
have set up a coordinate system with uniqueness on the principal axes of inertia of the bone in vivo.
The coordinate system is thus employed to determine the prime meridian, and the average radius of the
tomographic boundary is employed to determine the contour to map the bone surface. This means that
the bone surface mapping technique could present an alternate approach to study the bone’s morphology.
Standardized coordinate system of bone
The CT image of bone in vivo can generate the principal moments of inertia and the direction cosines [20].
Suppose the moment of inertia of the bone’s tissue relative to its center of mass is a constant; then
the magnitude of the bone’s principal moments of inertia will be determined by its shape and mass
distribution. The inertia tensor suggests that we can always find a group of coordinate systems where
2
three products of inertia will be nil at the same time [21, 22]. The magnitudes of these three principal
moments of inertia could come out with three results: (i) all three are equal; (ii) two out of three are
equal and (iii) each one is different from the other. When the object is homogeneous, in the first case, it
is a sphere; in the second, an ellipsoid, a cube, a cylinder or a rectangular. When the bone’s location and
orientation relative to inertia reference frame (i.e. the coordinate system is defined by the CT coordinate
system) are fixed, in the first case, there are numerous principal axes of inertia, while in the second
case, the orientation of one principal axis can be determined, but not the other two. Therefore, in the
first and second case, the principal moments of inertia have nothing to do with their principal axes of
inertia. In the third case, however, the direction cosines of the bone’s principal axis of inertia relative to
inertia reference frame are unique, which means a one-to-one corresponding relation between the direction
cosines of the bone’s principal axes of inertia and its shape.
If the bone is defined as a collection of elements of volume ∆V , the element’s position can be repre-
sented by (xoi, yoi, zoi) (where o is located in the center of mass ), then the magnitude of the principal
moments of inertia can be represented respectively by:
Ix = P(cid:0)y2
Iy = P(cid:0)x2
Iz = P(cid:0)x2
oi + z2
oi + z2
oi + y2
oi(cid:1) ρi∆V
oi(cid:1) ρi∆V
oi(cid:1) ρi∆V,
(1)
where ρ is the density, ∆V = ∆x∆y∆z, ∆x and ∆y are the pixel sizes and ∆z the layer distances of CT
images.
Let the angular displacements of the bone that take turns to rotate around axes x, y, z be α, β, γ.
According to Equations 1, we can set up the following equation:
(Iy − Iz)α = P[(yoi cos α − zoi sin α)2 − (yoi sin α + zoi cos α)2]ρi∆V
(Ix − Iz)β = P[(xoi cos β + zoi sin β)2 − (xoi sin β − zoi cos β)2]ρi∆V
(Ix − Iy)γ = P[(xoi cos γ − yoi sin γ)2 − (xoi sin γ + yoi cos γ)2]ρi∆V.
Next, differentiate Eq. 2, and let
d (Iy − Iz)α
dα
,
d (Ix − Iz)β
dβ
,
d (Ix − Iy)γ
dγ
,
which will generate the following equation:
α = 1
β = 1
γ = 1
2 arctan(cid:18)
2 arctan(cid:18)
2 arctan(cid:18)
2 P(yoizoiρi∆V )
oiρi∆V )−P(z2
P(y2
2 P(xαizαiρi∆V )
αiρi∆V )−P(z2
P(x2
2 P(xβiyβiρi∆V )
βiρi∆V )−P(y2
P(x2
oiρi∆V )(cid:19)
αiρi∆V )(cid:19)
βiρi∆V )(cid:19)
(A)
(B)
(C).
(2)
(3)
It is apparent that when and ONLY when Ix 6= Iy 6= Iz will Eq. 3 have a set of solutions. P yiziρi∆V ,
P ziziρi∆V and P xiyiρi∆V are three products of inertia of inertia tensor. Within the range of [0, π],
according to Eqs. 2 and 3, the limited rotations can always turn three products of inertia into zero at the
same time.
The bone’s shape is asymmetrical and its structure is anisotropic [23–27]. Eq. 3 exposes the direction
cosines of principal axes of inertia relative to the inertia reference frame of the bone characterization are
unique. As a result, the coordinate system set upon the principal moment of inertia can not only depict
the position and orientation of the bone, but also verify the bone surface shape and its changes when
making a quantitative analysis. Eqs. 2 and 3 also suggest that we can set up a coordinate system whose
coordinate origin is located arbitrarily at the center of mass of the bone. After limited rotations, the
coordinate axes is positioned on the principal axes of inertia.
Mapping of the Bone surface
CT scanning simplifies the bone as a collection of elements of volume ∆V (different densities). When
performing an isotropic scanning (where pixel size must be the same as the layer distance), the volume
of ∆V is a constant. The position of ∆V relative to the center of mass differs from one another. When
the result of Eq. 3 is replaced for that in Eq. 2 accordingly and when the bone’s coordinate system is
positioned on the principal axes of inertia, the rotation will change the original CT image. It is necessary,
then, to reconstruct the new image, which can be performed by the following equation:
3
∆x
xi = trunc(cid:16) xoi−min(xoi)
yi = trunc(cid:16) yoi−min(yoi)
zi = trunc(cid:16) zoi−min(zoi)
(cid:17) ∆x + min(xoi)
(cid:17) ∆y + min(yoi)
(cid:17) ∆z + min(zoi),
∆y
∆z
(4)
where (xoi, yoi, zoi) stands for the position of ∆V after rotation, (xi, yi, zi) for that of the reconstructed
tomogram and trunc() for a function that truncates a number to an integer by removing the fractional
part of the number. To keep the CT images isotropic, ∆dx = ∆dy = ∆dz is defined in Eq. 4, and its
pixal size and layer distance are kept the same of those of the original image. According to Eq. 4, when
keeping the ∆V to be in a cube, the new CT image after rotation remains to be closed and continuous.
CT scanning divides the bone surface into a collection of the tomographic image boundaries. In this
way, the mapping of the bone surface has become an issue to develop the tomographic boundary, making
it necessary to detect and draw the tomographic boundary. Before a new CT scanning, the equipment is
reset, i.e. the gray value of the air is set as zero. The scanned tomographic images of bone are processed
by Eq. 4, and their boundaries are drawn by the following equation:
ρ(x, y)z =
ρ(x, y)z
ρ(x, y)z
ρ(x, y)z
ρ(x, y)z
0
other,
ρ(x, y)z > 0, ρ(x + 1, y)z = 0
ρ(x, y)z > 0, ρ(x − 1, y)z = 0
ρ(x, y)z > 0, ρ(x, y + 1)z = 0
ρ(x, y)z > 0, ρ(x, y − 1)z = 0
(5)
where z is the number of sequence of tomogram, (x, y)z the position of ∆V relative to the tomographic
center of mass and ρ(x, y)z the density of (x, y)z.
When mapping the bone surface, the bone will be ”cut” - from a cylindrical surface to a rectangle,
or a rhombus. However it is cut, its ultimate area would be the same. But when it is a rectangle, there
is only one. How to cut the bone into a rectangle? The two principal axes of inertia (the minimal and
maximal principal moment of inertia) form a plane. On this plane, the bone surface boundary is called
the prime meridian, which is used as the surface cutting line to develop the bone surface. The following
equation will make it happen:
i =
z > yc
z
xi
z = xc
z − xc
xi
xi
z − xc
z, yi
z = −∆x, yi
z = −2∆x, yi
0
1
2
· · ·
z − xc
xi
n − 1
z = ∆x, yi
xi
z − xc
n
z = 2∆x, yi
z > yc
z
z > yc
z,
z > yc
z
z > yc
z
(6)
where (xi
z, yi
of mass, (xc
the sequence of ∆V of the boundary after being cut.
z) is the position of the ∆V at the tomographic boundary relative to the tomographic center
z, yc
z) the position of tomographic center of mass relative to the inertia reference frame and i
Eq. 6 sequences the ∆V s at the tomographic boundary which has been cut. The average radium of
the tomographic boundary perpendicular to the minimal (maximal) principal moment of inertia is defined
as the ”sea level”, which is used as a datum line so that the tomographic boundary can be developed by
the following equation:
4
p(x, y)z = p(cid:0)i + xc
z, hi
z(cid:1)z ,
z = ri
z − rz, ri
(7)
z − xc
z)2 + (yi
z − yc
z)2 and
where z shares the same definition of that in Eq. 5, hi
rz = P rj
z) and (xi
z, yc
z, yi
n , (xc
z
z) have the same definitions as those in Eq. 6.
z = p(xi
Eqs. 4- 7 have developed the closed surface into an open three-dimensional curved one with the
properties of a contour. The three-dimensional curved map of the bone surface can be further developed
into a two-dimensional plane. We can, however, translate the bone surface into a plane with the help of
the following equation:
p(x, y) = p(cid:18)Z (cid:18)ql2(i) + Z 2
i(cid:19) di,Z (cid:16)ql2(j) + Z 2
j(cid:17) dj(cid:19) ,
(8)
where p(x, y) presents the position of ∆V in a plane whose surface has been flattened, and
l(i) = q(zi − zi−1)2 + Z 2
i , Zi = Z zi − zi−1di,
l(j) = q(zj − zj−1)2 + Z 2
j , Zj = Z zj − zj−1dj,
z stands for the value of contour on (i, j) in bone surface mapping.
Eq. 8 suggests that the mapping of the bone surface actually serves as a simulation of the bone surface
structure. It is a space model of an image symbol to represent the bone surface. It shares the consistency
with the real body of the bone surface structure.
Experiment
Prior to our study, the Ethic Committee of Guangzhou Institute of Physical Education has proved our
study and the participant has provided fully informed consent to participate in this study by signing a
written consent form. From January 2008 to August 2009, we followed the track of Guangdong Provincial
Youth Team of Male Wrestlers by using a 64 slice scanner (Brilliance 64, Philips Medical Systems).
Excluding team members who left the team halfway and those with injuries, in January 2008 and August
2009, we were able to scan and collect the data of a 25-year-old wrestler’s sesamoid bones beneath the
head of the first metatarsal bone of both feet. See Table 1 for information of the sesamoid bones.
In Table 1, the volumes of left and right foot’s internal and external sesamoid bone have changed by
0.26%, −1.98%, 0.38% and −1.32% respectively; their surface areas have changed by 0.22%, −0.41%,
0.54% and −0.23% respectively and their density by 0.85%, 1.45%, 2.40% and 3.77% respectively.
Let’s make a mapping analysis to the right foot’s external sesamoid beneath the head of the first
metatarsal bone. First, rotate the first-time scanned sesamoid bone around axis x from 0 to 180 degrees.
According to Eq. 2, the variations of (Iy − Iz)α, (Ix − Iz)β and (Ix − Iy)γ are shown in Fig. 1, indicating
that within the range of rotation from 0 to 180 degrees, an extremum exists in Eq. 2. We can get the
result of α = 157.73 when the extremum is calculated by Eq. 3A. After the sesamoid bone rotates around
axis x at 157.63 degree, it then rotates around axis y. The calculation by Eq. 3B generates the result
of β = 8.92. When the sesamoid bone rotates around axis y at 8.92 degree, the calculation of Eq. 3C
generates the result of γ = 172.07. When the coordinate system of the sesamoid bone rotate around axes
xyz at 157.63/8.92/172.07 degree respectively, the axes of of the sesamoid bone coincide with those of
the principal moments of inertia.
It’s shown that the morphologically asymmetric and heterogeneously distributed bone has the unique-
ness of direction cosines of their principal moments of inertia. An arbitrary coordinate system set upon
the bone’s center of mass can make every coordinate axis coincide with the principal moment of inertia by
5
coordinate transformation, i.e. the proposed principal axis’s coordinate system has its uniqueness. This
method can be applied to both the homogeneous asymmetric geometry and the heterogeneous asymmetric
geometry.
The principal axes of sesamoid bone is set up by Eqs. 2 and 3; the tomography of sesamoid bone is
reconstructed after it is rotated by Eq. 4; the boundary of the sesamoid bone is drawn by Eq. 5. When
the prime meridian is determined by the principal axes, the sesamoid bone surface is developed by Eq. 6
and the mapping of the sesamoid bone surface is accomplished by Eq. 7. See Fig. 2.
Figs. 2A and 2B indicate that the mapping technique can better illustrate the bone surface properties.
Fig. 2C suggests that the mapping technique can be used as a quantitative method to study the changes
of an object’s shape. Using the bone surface mapping, Eq. 8 can flatten the bone surface as a plane. See
Fig. 3.
It can be concluded that the professional training has caused adaptative changes of the sesamoid
bone’s shape. The structural changes can be depicted by the density and distribution while the shape
and its changes can be analyzed by the bone surface mapping.
Conclusion
The generalized Papoulis theorem [29] elucidates that the bone surface shape keeps its geometric invari-
ance, such as rotation, translation or dimension change [30, 31]. This ensures the consistency of the CT
scanning results of different postures of bone in vivo when its isotropy is ascertained. The uniqueness
of the relative consistency of the inertia reference system of principal moments of inertia on direction
cosines provide evidence to the bone surface mapping technique. The characters such as the geometric
invariance and the uniqueness of the coordinate system of the principal moments of inertia have enabled
the bone surface mapping technique to depict the bone’s external morphological characters. This can
advance the research of the morphological mechanisms. The experiment of the bone in vivo signifies that
the bone mapping technique adds another research method and supplements the analytical method of
the bone’s three-dimensional imaging technique.
We can understand the world through a map. When the bone surface mapping technique reveals
its surface information through a ”map”, the activities of our life evolve continuously on this map. We
anticipate that this mapping technique will be widely used in related disciplines.
Acknowledgments
The authors would like to acknowledge the support from the subject.
References
1. Buckwalter JA, Glimcher MJ, Cooper RR, Recker R (1995) Bone biology. II: Formation, form,
modeling, remodeling, and regulation of cell function. J Bone Joint Surg Am 77: 1276-1289.
2. Nomura S, Takano-Yamamoto T (2000) Molecular events caused by mechanical stress in bone.
Matrix Biol 19: 91-96.
3. Hugate R, Pennypacker J, Saunders M, Juliano P (2004) The effects of intratendinous and retro-
calcaneal intrabursal injections of corticosteroid on the biomechanical properties of rabbit Achilles
tendons. J Bone Joint Surg Am 86:794-801.
4. Canalis E, Giustina A, Bilezikian JP (2007) Mechanisms of anabolic therapies for osteoporosis. N
Engl J Med 357: 905-916.
6
5. Anas A, Arnott R, Small KA (1998) Urban spatial structure. J Econ Lit 36: 1426-1464.
6. Legendre P (1993) Spatial autocorrelation: trouble or new paradigm? Ecology 74: 1659-1673.
7. Barnes WL, Dereux A, Ebbesen TW (2003) Surface plasmon subwavelength optics. Nature 424:
824-830.
8. Wang BC (1985) Resolution of phase ambiguity in macromolecular crystallography. Methods En-
zymol. 115: 90-112
9. Suri J, Singh S, Reden L (2002) Computer vision and pattern recognition techniques for 2-D and
3-D MR cerebral cortical segmentation (Part I): A state-of-the-art review. Pattern Anal Appl 5:
46-76.
10. Cuntz H, Forstner F, Haag J, Borst A (2008) The morphological identity of insect dendrites. PLoS
Comput Biol 4(12): e1000251.
11. Bates E, Wilson SM, Saygin AP, Dick F, Sereno MI, Knight RT, et al. (2003) Voxel-based lesion-
symptom mapping. Nat Neurosci 6: 448-50.
12. Morley M, Molony CM, Weber TM, Devlin JL, Ewens KG, et al. (2004) Genetic analysis of genome-
wide variation in human gene expression. Nature 430: 743-747.
13. Saenz M, Buracas GT, Boynton GM (2002) Global effects of feature-based attention in human
visual cortex. Nat Neurosci 5: 631-632.
14. Bennis C, Vezien JM, Iglesias G (1991) Piecewise flattening for non-distorted texture mapping.
ACM SIGGRAPH Computer Graphics. 25: 237-246.
15. Haker S, Angenent S, Tannenbaum A, Kikinis R, Sapiro G, Halle M (2000) Conformal surface
parameterization for texture mapping. IEEE TVCG 6: 181-189.
16. Keyak JH, Meagher JM, Skinner HB, Mote CD (1990) Automated three-dimensional finite element
modelling of bone: a new method. J Biomed Eng 12: 389-397.
17. Bouxsein ML, Karasik D (2006) Bone geometry and skeletal fragility. Curr Osteoporos Rep 4:
49-56.
18. Matthews F, Messmer P, Raikov V, Wanner GA (2009) Ptient-specific three-dimensional composite
bone models for teaching and operation planning. J Digit Imaging 22: 473-482.
19. Davies ME, Abalakin VK, Bursa M, Lieske, J H, Morando B, Morrison D, Seidelmann PK, Sinclair
AT, Yallop B, Tjuflin YS (1996) Report of the IAU/IAG/COSPAR Working Group on cartographic
coordinates and rotational elements of the planets and satellites: 1994. Celest Mech Dynam Astron
63: 127-148.
20. Coburn JC, Upal MA, Crisco JJ (2007) Coordinate systems for the carpal bones of the wrist. J
Biomech 40: 203-209.
21. Hinrichs RN, Lallement SR, Belson RC (1982) An investigation of the inertial properties of back-
packs loaded in various configurations. Tech Rep NATICK/TR-82/023 1: 1-74.
22. Nikravesh P, Lin YS (2005) Use of principal axes as the floating reference frame for a moving
deformable body. Multibody Syst Dyn 13: 211-231.
7
23. Biewener AA, Thomason J, Lanyon LE (1983) Mechanics of locomotion and jumping in the forelimb
of the horse (Equus): in vivo stress developed in the radius and metacarpus. J Zool 201: 67-82.
24. Hans D, Fuerst T, Duboeuf F (1997) Quantitative ultrasound bone measurement. Eur Radiol 7
(Suppl.2): S43-S50.
25. Kleber AG, Rudy Y (2004) Basic mechanisms of cardiac impulse propagation and associated ar-
rhythmias. Physiol Rev 84: 431-488.
26. Ketcham RA, Ryan TM (2004) Quantification and visualization of anisotropy in trabecular bone.
J Microscopy 213: 158-171.
27. Fink WL, Humphries JH (2010) Morphological description of the extinct North American Sucker
Moxostoma lacerum (Ostariophysi, Catostomidae), Based on high-resolution X-Ray computed to-
mography. Copeia 2010: 5-13.
28. Gu X, Gortler SJ, Hoppe H, (2002) Gometry images. ACM SIGGRAPH 355-361
29. Papoulis A, Signal Analysis. New-York: McGraw-Hill, 1977.
30. Nguyen N, Milanfar P (2000) A wavelet-based interpolation-restoration method for superresolution
(wavelet superresolution). Circ Syst Signal Pr 19: 321-338.
31. Rochefort G, Champagnat F, Le Besnerais G, Giovannelli JF (2006) An improved observation
model for super-resolution under affine motion. IEEE T Image Process 15: 3325-3337.
Table 1: Sesamoid bone volume, surface area and density of two measurements
8
First
volume
mm3
307.06
204.35
257.84
205.87
surface
mm2
247.41
184.18
209.59
181.61
Left - external
Left - internal
Right - external
Right - internal
Second
volume
density
mg/mm3 mm3
1.84
1.94
1.87
1.89
307.87
200.38
28.83
203.13
surface
mm2
247.96
183.43
210.73
181.20
density
mg/mm3
1.85
1.97
1.92
1.97
Iy−Iz
Ix−Iz
Ix−Iy
i
s
e
x
a
l
a
p
c
n
i
r
p
e
h
t
f
o
a
i
t
r
e
n
i
f
o
s
t
n
e
m
o
M
2000
1500
1000
500
0
−500
−1000
−1500
−2000
0
50
100
150
Angular displacement
200
250
300
350
Figure 1: The variations of the axis moment of inertia of the first measurement of the wrestler’s lower
external sesamoid bone of the first metatarsal bone accompanying the changes of right foot.
9
A
5
0
e
n
i
l
r
u
o
t
n
o
C
−5
100
B
5
0
e
n
i
l
r
u
o
t
n
o
C
−5
100
80
60
40
20
Principal axis of inertia (min)
0
0
20
100
80
80
60
60
40
Principal axis of inertia (max)
Principal axis of inertia (min)
40
20
0
0
20
60
40
Principal axis of inertia (max)
100
80
C
5
0
e
n
i
l
r
u
o
t
n
o
C
−5
100
80
60
40
Principal axis of inertia (min)
20
0
0
20
100
80
60
40
Principal axis of inertia (max)
Figure 2: Surface mapping of bone. Fig. 2A Surface mapping of the first measurement of the wrestler right
foot’s external sesamoid bone beneath the head of the first metatarsal bone. Fig. 2B Surface mapping of
the second measurement of the wrestler right foot’s external sesamoid bone beneath the head of the first
metatarsal bone. Fig. 2C Variations of the surface mapping of the wrestler right foot’s external sesamoid
bone beneath the head of the first metatarsal bone after 18 months (the second measurement). Figs. 2A,
2B and 2C have been rated by percentage and smoothed. A closed bone surface means that when it
is continuous, there is no boundary. But when the surface is cut by the prime meridian, a boundary
emerges. So when smoothing the surface, cloud computing method is adopted [28] to ensure the integrity
of the object’s shape.
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
)
x
a
m
(
a
i
t
r
e
n
i
f
o
i
s
x
a
l
i
a
p
c
n
i
r
P
0
0
0.1
0.2
A
B
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
)
x
a
m
(
a
i
t
r
e
n
i
f
o
i
s
x
a
l
i
a
p
c
n
i
r
P
0.3
0.4
0.5
0.6
Principal axis of inertia (min)
0.7
0.8
0.9
1
0
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Principal axis of inertia (min)
Figure 3: Flattened plane of bone. Fig. 3A The flattened bone surface of the first measurement of the
wrestler’s lower external sesamoid bone of the first metatarsal bone. Fig. 3B The flattened bone surface
of the second measurement of the wrestler’s lower external sesamoid bone of the first metatarsal bone
after 18 months.
|
1812.02296 | 1 | 1812 | 2018-12-04T10:19:06 | Temperature changes accompanying signal propagation in axons | [
"physics.bio-ph",
"physics.med-ph"
] | In this paper mathematical models are formulated in order to simulate heat production and corresponding temperature changes which accompany the propagation of an axon potential. Based on earlier experimental results, several models are proposed. Together with the earlier system of coupled differential equations derived by the authors for describing the electrical and mechanical components of signalling in nerve fibres, the novel results permit to cast the whole process of signalling into one system. The emphasis is on the mathematical description of coupling forces. The numerical results are qualitatively similar to experiments. | physics.bio-ph | physics |
Temperature changes accompanying signal propagation in axons
Kert Tamm∗, Juri Engelbrecht, Tanel Peets
Department of Cybernetics, School of Science, Tallinn University of Technology, Akadeemia tee 21, Tallinn 12618, Estonia
E-mail: [email protected], [email protected], [email protected]
Abstract. In this paper mathematical models are formulated in order to simulate heat production and corresponding temperature
changes which accompany the propagation of an axon potential. Based on earlier experimental results, several models are proposed.
Together with the earlier system of coupled differential equations derived by the authors for describing the electrical and mechanical
components of signalling in nerve fibres, the novel results permit to cast the whole process of signalling into one system. The
emphasis is on the mathematical description of coupling forces. The numerical results are qualitatively similar to experiments.
Key words: mathematical modelling, action potential, heat production, temperature.
1.
INTRODUCTION
The generation of heat accompanying the propagation of an action potential has been measured in many
experimental studies by Abbott et al [1], by Downing et al [3], by Howarth et al [11], by Ritchie et al [15],
by Tasaki et al [16, 17] etc. However, there is no generally accepted mechanism for heat production and
temperature changes because the measurements differ in a large scale. The generally accepted energy of a
membrane capacitor is related to the square of voltage [15]. Tasaki and Byrne [17] have shown that either
the generated temperature or its rate may be related to voltage depending on physical properties. It is of great
interest to formulate the mathematical basis for calculating the heat production or temperature changes given
these experimental assumptions. In our earlier studies [4 -- 7, 14] a system of differential equations is derived
in order to model the action potential (AP) and accompanying mechanical effects -- pressure wave (PW) in
axoplasm, longitudinal (LW) and transverse (TW) waves in the surrounding biomembrane. This system will
here be enlarged by including the temperature (Θ) effects. The model equations are presented with the focus
on possible models of heat and temperature production in Section 2. Further on, the numerical examples are
given in dimensionless setting in Section 3. The final remarks are presented in Section 4.
2. MODEL EQUATIONS
The coupled system of model equations has been previously proposed by the authors for describing
the nerve pulse (action potential (AP)) propagation including the accompanying mechanical effects (pres-
sure wave (PW) and longitudinal density change in the lipide bi-layer (LW)) [4 -- 7, 14]. In this paper, the
temperature effects which accompany the propagation of an AP in the axoplasm, are also described. The
dimensionless 1D model used so far [4 -- 7, 14] is the following:
ZT = DZX X + Z(cid:0)Z − [a1 + b1] − Z2 + [a1 + b1] Z(cid:1) − J,
JT = ε([a2 + b2] Z − J) ,
UT T = c2UX X + NUUX X + MU 2UX X + NU 2
X + 2MUU 2
X − H1UX X X X + H2UX X T T + F1(Z , J, P),
PT T = c2
f PX X − µPT + F2(Z , J,U ),
(1)
(2)
(3)
where Eq. (1) is the FitzHugh -- Nagumo (FHN) equation [13], Eq. (2) is the improved Heimburg -- Jackson
(iHJ) equation [6, 8, 9], Eq. (3) is the classical wave equation with the added viscous dampening term.
In Eq. (1) Z is potential, J is the abstracted ion current (like characteristic to the FHN equation), ai , bi
are "electrical" and "mechanical" activation coefficients, D,ε are coefficients. In Eq. (2) U is longitudinal
∗ Corresponding author
2
density change, c is sound velocity of unperturbed state in the lipid bi-layer, N , M are nonlinear coefficients,
H1 , H2 are dispersion coefficients. In Eq. (3) P is pressure, c f is sound velocity in axoplasm, µ is dampening
coefficient. Here and further index X denotes spatial partial derivative and index T denotes partial derivative
with respect to time.
Coupling terms have been described by polynomials
F1 = γ1
PT
1 + U
+ γ2
JT
1 + U
− γ3
ZT
1 + U
; F2 = η1ZX + η2JT + η3ZT ,
(4)
where γi are coupling coefficients for mechanical wave and ηi are coupling coefficients for the pressure
wave. Time derivatives act across membrane and spatial derivatives act along the axon.
What has not been considered in the previous publications on mathematical modelling is the possible
temperature change accompanying the AP. As far as system (1) -- (3) is a system of partial differential
equations (PDE), we look for an additional equation capable to model the temperature changes. The cru-
cial question is how temperature effects are related to other propagating effects and as before we look for
corresponding coupling forces which in this case should be related to the heat production. It is worth also
keeping in mind that temperature change normally is not considered like a wave but is considered to be
like a diffusive process. There are several ideas proposed in earlier studies on thermal changes caused by
propagating APs. It has been stated that the energy of the membrane capacitor Ec is [15]
Ec =
1
2
CmZ2
,
(5)
where Cm is a capacitance and Z is the amplitude of the AP. Noting that heat energy Q ≈ Ec, it can be
deduced that Q should be proportional to Z2. The standard Fourier law (the thermal conductivity equation)
in its simplest 1D form is
Q = −kΘX ,
(6)
where Q is the heat energy, k is the thermal conductivity and Θ is the temperature. Note that heat energy is
proportional to the negative temperature gradient. Combining (5) and (6), we obtain
Θ ∝
Cm
2k Z Z2dX .
(7)
Further we follow experimental results by Abbot et al [1] which demonstrate that the heat increase at the
surface of the fibre is positive. It has also been argued that either Z ∝ Θ or Z ∝ ΘT depending on physical
properties [17]. The potential Z of the AP or its square Z2 might be not the only sources. For example, in
[10, 12] it is argued in favour of the idea that experimentally observed temperature changes might be the
result of the propagating mechanical wave in the lipid bi-layer. As there seems to be no clear consensus
which quantities associated with the nerve pulse propagation are the sources of the thermal energy, we will
consider here all the possibilities for the coupling of the waves in the ensemble to the additional model
equation for the temperature.
The idea is to cast these ideas into a mathematical form. As far as temperature is a function of space and
time we opt to use the classical heat equation which is a parabolic PDE describing the distribution of heat
(or variation in temperature) in a given region over time. It is straightforward [2] to derive the heat equation
in terms of temperature from the Fourier's law by considering the conservation of energy
where α is the thermal diffusivity. In our case, a possible source term must be added to Eq. (8)
ΘT = αΘX X + F3(Z , J,U ).
ΘT = αΘX X ,
(8)
(9)
The source term F3 permits to account for different assumptions proposed to describe the temperature gener-
ation and consumption. Further on, a number of different functions F3 are used for this purpose: (i) F3 = τ1Z
and F3 = τ2Z2 (see Fig. 2); (ii) F3 = τ3J and F3 = τ4J2 (see Fig. 3); (iii) F3 = τ5U and F3 = τ6U 2 (see Fig. 4);
(iv) F3 = τ7ZT + τ8JT and F3 = τ9JT + τ10UX , (see Fig. 5) where τi are the thermal coupling coefficients.
3
3. NUMERICAL EXAMPLES AND DISCUSSION
The model equations are solved using the pseudospectral method (PSM) [7]. The added heat equation
(9) can be directly solved as outlined in [7] following the same procedure without further modifications.
The idea of the PSM is write the PDEs in a form where all the time derivatives are on left hand side (LHS)
and all the spatial derivatives on the right hand side (RHS) of the equation and then apply properties of the
Fourier transform for finding the spatial derivatives reducing the PDE into an ordinary differential equation
(ODE) which can be solved by standard numerical methods. For the numerical solving -- a localized initial
condition is used in the middle of the spatial period for Z with the rest of the processes assumed to be at
rest (zero initial conditions). This initial "spark" is taken above the threshold value in the middle of spatial
period causes the AP to emerge and propagate to the left and right. All other waves are generated by the
propagating AP as a result of coupling terms added to the system. Boundary conditions are taken periodic as
is required for the PSM, moreover in the present paper the time integration intervals have been taken short
enough to avoid interaction between counter-propagating waves at the boundaries. In the figures only left
propagating waves are plotted.
The following parameter values have been used for the numerical simulations:
(i) for the FHN equation (1)
D = 1;ε = 0.018; a1 = 0.2; a2 = 0.2; b1 = −0.05 ·U ; b2 = −0.05 ·U ,
(ii) for the improved Heimburg-Jackson model (2)
c2 = 0.10; N = −0.05; M = 0.02; H1 = 0.2; H2 = 0.99;
(iii) for the pressure (3)
c2
f = 0.09; µ = 0.05;
(iv) for the heat equation (9)
α = 0.05;
(v) and the coupling coefficients are
γ1 = 0.008;γ2 = 0.01;γ3 = 3 · 10−5;η1 = 0.005;η2 = 0.01;η3 = 0.003;
τ1,...,6 = 5 · 10−5;τ7 = 5 · 10−5;τ8 = 1 · 10−3;τ9 = 1 · 10−3;τ10 = 5 · 10−5
.
The parameters for the numerical scheme are L = 64π (the length of the spatial period), Tf = 400 (the end
time for integration, figures in the present paper are shown at T = 400), n = 2048 (the number of spatial
grid points), the initial amplitude for Z is Az = 1.2 and width of the initial sech2 type pulse is Bo = 1.0 [7].
1
0.8
0.6
0.4
0.2
0
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
40
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
0.8
0.6
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
120
140
40
80
60
X - dimensionless space
100
80
60
X - dimensionless space
100
120
140
Fig. 1. The wave ensemble modelled quantities (left) and the TW derived from the LW (right).
The wave ensembles are shown in Fig. 1. It should be noted that some characteristics normally observed
in experiments, like the transverse displacement (TW) of the membrane is not described by a differential
equation but can be derived directly from the calculated LW as a derivative [6]. Certain oscillations of the
4
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
1
0.8
0.6
0.4
0.2
0
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
120
140
40
40
80
60
X - dimensionless space
100
80
60
X - dimensionless space
100
120
140
Fig. 2. The temperature change Θ and thermal energy change Q = ΘX if F3 = τ1Z (left) and if F3 = τ2Z2(right).
1
0.8
0.6
0.4
0.2
0
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
120
140
40
40
80
60
X - dimensionless space
100
80
60
X - dimensionless space
100
120
140
Fig. 3. The temperature change Θ and thermal energy change Q = ΘX if F3 = τ3J (left) and if F3 = τ4J2(right).
calculated TW reflect the physical properties of the biomembrane which is described by a Boussinesq-type
equation [6].
Few more remarks are in order. Note that in Figs 2 -- 5 thermal energy change Q is plotted as Q = ΘX .
One can observe in Fig. 2 that in the case of temperature increase proportional to Z the local temperature
drops in the negative polarity region of the AP while if the temperature increases proportional to Z2 the local
temperature keeps increasing even in the negative polarity region of the AP. Observing correlation between
the Z (or Z2) and the measured temperature increase near axon seems to be common for several experiments
by Howarth et al [11], Ritchie et al [15] and by Tasaki et al [17]. Experimental results published so far
seem inconclusive to argue firmly in favour of either option as normally only some kind of averaged thermal
energy production over some time can be measured. Another note made is that while experimentally the heat
production proportional to AP has been observed, this correlation does not have to mean direct causality.
Like generally in nature, the potential gradient alone is rarely the only source of the temperature change
in an environment. One could argue that heat production might be instead be proportional to ion currents
(Fig. 3) or the longitudinal density change (Fig. 4) which accompany the propagating AP. For the sake of
completeness temperature increase as a function of J and J2 is presented in Fig. 3 and as a function of
density change U and U 2 in Fig. 4. The idea behind including J and U as sources is that mechanisms where
the temperature increase is caused by the current flowing through the environment or by the deformation of
the solid are well established in the physics even if these are not as common sources assumed for the heat
1
0.8
0.6
0.4
0.2
0
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
0.8
0.6
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
1
0.8
0.6
0.4
0.2
0
-0.2
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
120
140
40
40
80
60
X - dimensionless space
100
80
60
X - dimensionless space
100
5
120
140
Fig. 4. The temperature change Θ and thermal energy change Q = ΘX if F3 = τ5U (left) and if F3 = τ6U 2(right).
s
e
u
l
a
v
d
e
z
i
l
a
m
r
o
N
0.8
0.6
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
120
140
40
40
80
60
X - dimensionless space
100
80
60
X - dimensionless space
100
120
140
Fig. 5. The temperature change Θ and thermal energy change Q = ΘX if F3 = τ7ZT + τ8JT (left) and if F3 = τ9JT + τ10UX (right).
generation [10, 12] as the correlation with the AP [11, 15, 17].
In such a context it might not be a problem to observe that some heat energy "goes away" in the negative
polarity phase of the used driving signal if one assumes heat production proportional to the driving signal
as the underlying endothermic mechanism might be actually independent of the AP propagation, like, some
kind of endothermic chemical reaction starting [1]. The open question in the latter case is, however, the
issue of time scales. Moreover, another effect present in the numerical simulation results presented must
be noted -- at least in part the reason why the temperature can be lower in the areas through which the
signal has already propagated as opposed to the front is that the driving signal can be actually of lower
amplitude at earlier stages of the signal propagation. This is the main reason why in Fig. 4 there is quite
a significant drop in temperature for the case of U 2 (right panel) because in this case the change in driving
signal amplitude is amplified in time. The signal shape in Fig. 4 (right panel) is qualitatively similar to the
signal shapes observed by Abbott, et al in [1]. Another note relevant specifically for the Eq. (2) can be made
-- it is a conservative equation before adding the coupling term F1 facilitating energy exchange between
the coupled equations. However, if the mechanical wave in the lipid bi-layer is considered as a source for
thermal energy in the system for earnest this equation would need some kind of additional term which would
take some energy away from the mechanical wave. For example, the simplest possibility might be using a
Voigt -- type model which means appearing a term UX X X in Eq. (2).
Experimentally the thermal energy production and consumption has been observed in the time scales
6
comparable to the AP propagation [16, 17] to few orders of magnitude longer as shown by Howarth et al
[11]. Following the idea that the underlying processes responsible for the heat production and consumption
might be proportional to some of the processes we are actually modelling one possibility shown in Fig. 5
where time derivatives of AP and ion current J (left) or the ion current time derivative and the gradient of
longitudinal density change (right) are used for the thermal energy generation and consumption. In such a
case thermal energy is generated when the driving signal is growing and consumed when the driving signal
is decreasing. The examples shown in Fig. 5 are intended to highlight what is mathematically possible
in the present framework and not as an implication of the thermal response being actually generated and
even more importantly, consumed, by the used signals in the reality. However some of the experimental
observations are qualitatively similar to profiles shown in Fig. 5 [17]. A minor remark on whether to use
spatial or temporal derivatives and signals is in order -- in our 1D setting the signal shape is practically the
same regardless if observed in space (at fixed time) or in time (at fixed spatial point). Actually, integrating
Eq. (9) in time with F3 ∝ Z2 (Fig. 2, right panel) takes into account expression (7).
Finally it seems prudent to emphasize where, exactly, is this temperature change modelled on axon in
the presented framework as normally the experimental measurements for the phenomenon are performed
close to the axon wall outside of the axon itself. The temperature changes in Figs 2, 3 and 5 (left panel) are
for the axoplasm inside the axon. The temperature plotted in Fig. 4 would have to be inside the lipid bi-layer
forming the axon wall because that is where the density change U is taken into account in our model which
is used to generate these changes in Fig. 4. Overall, it is assumed that any temperature changes measured
sufficiently close to the lipid bi-layer forming the axon wall are proportional to the temperature changes in
the axoplasm or in the lipid bi-layer giving us, in essence, a combined reading of all available temperature
changes in close proximity of the hypothetical measuring point (which is the case for Fig. 5 right panel). In
the present framework a possible time delay between heat being generated or consumed and the "measured"
signal a small distance away is not taken into account.
One possible extension for the system (1) -- (3), (9) is to include a separate PDE into the system for
capturing the temperature decrease in a manner which better aligns with some of the experimental results.
For example, the temperature response can happen on a noticeably larger time scale than the AP propagation,
implying potentially a different underlying mechanism as described by Abbott et al [1] and Howarth et al
[11].
4. FINAL REMARKS
An attempt is made to cast the experimental ideas [11, 15 -- 17] into the mathematical form. The ex-
perimental results differ on a large scale -- the temperature profile follows the AP [16] and the temperature
changes are on much longer scale compared to the AP [1]. In this paper it is demonstrated that despite of
such scattered results, it is possible to construct a mathematical model reproducing the experimental results.
It seems that it is possible mathematically simulate the temperature generation as measured by Tasaki [16]
for the garfish olfactory nerve as well as the heat production measured by Howarth et al [11] for mammalian
nerves. However, the essence of coupling terms needs experimental verification. The temperature changes
are governed by diffusion contrary to the wave-like behaviour of a signal in a fibre. The mathematical mod-
els described above form certainly just a basic platform for further refinements of modelling. For example,
the effects of various ion currents or general oxygen consumption, CO2 output, the influence of carbohy-
drates (see [3]) should also be taken into account. But even within the present model, results described
above may enlarge the fundamental thermodynamics of nervous signals [10].
ACKNOWLEDGEMENTS
This research was supported by the European Union through the European Regional Development Fund
(Estonian Programme TK 124) and by the Estonian Research Council (projects IUT 33-24, PUT 434).
7
References
1. Abbott, B. C., Hill, A. V., and Howarth, J. V. The positive and negative heat production associated with
a nerve impulse. Proc. R. Soc. B Biol. Sci., 1958, 148(931), 149 -- 187.
2. Carslaw, H. and Jaeger, J. Conduction of heat in solids. Oxford Science Publications, Oxford, 1959.
3. Downing, A. C., Gerard, R. W., and Hill, A. V. The heat production of nerve. Proc. R. Soc. B Biol. Sci.,
1926, 100(702), 223 -- 251.
4. Engelbrecht, J., Peets, T., and Tamm, K. Electromechanical coupling of waves in nerve fibres. Biomech.
Model. Mechanobiol., 2018, 17(6), 1771 -- 1783.
5. Engelbrecht, J., Peets, T., Tamm, K., Laasmaa, M., and Vendelin, M. On the complexity of signal
propagation in nerve fibres. Proc. Estonian Acad. Sci., 2018, 67(1), 28 -- 38.
6. Engelbrecht, J., Tamm, K., and Peets, T. On mathematical modelling of solitary pulses in cylindrical
biomembranes. Biomech. Model. Mechanobiol., 2015, 14(1), 159 -- 167.
7. Engelbrecht, J., Tamm, K., and Peets, T. Modeling of complex signals in nerve fibers. Med. Hypotheses,
2018, 120, 90 -- 95.
8. Heimburg, T. and Jackson, A. On the action potential as a propagating density pulse and the role of
anesthetics. Biophys. Rev. Lett., 2007, 02(01), 57 -- 78.
9. Heimburg, T. and Jackson, A. D. On soliton propagation in biomembranes and nerves. Proc. Natl. Acad.
Sci. USA, 2005, 102(28), 9790 -- 9795.
10. Heimburg, T. and Jackson, A. D. Thermodynamics of the nervous impulse. In Struct. Dyn. Membr.
interfaces (Kaushik, N., ed.). John Wiley & Sons, 2008, 318 -- 337.
11. Howarth, J. V., Keynes, R. D., and Ritchie, J. M. The origin of the initial heat associated with a single
impulse in mammalian non-myelinated nerve fibres. J. Physiol., 1968, 194(3), 745 -- 93.
12. Mussel, M. and Schneider, M. F. It sounds like an action potential: unification of electrical, chemical
and mechanical aspects of acoustic pulses in lipids. 2018. http://arxiv.org/abs/1806.08551.
13. Nagumo, J., Arimoto, S., and Yoshizawa, S. An active pulse transmission line simulating nerve axon.
Proc. IRE, 1962, 50(10), 2061 -- 2070.
14. Peets, T. and Tamm, K. On mechanical aspects of nerve pulse propagation and the Boussinesq paradigm.
Proc. Est. Acad. Sci., 2015, 64(3), 331.
15. Ritchie, J. M. and Keynes, R. D. The production and absorption of heat associated with electrical activity
in nerve and electric organ. Q. Rev. Biophys., 1985, 18(04), 451.
16. Tasaki, I. A macromolecular approach to excitation phenomena: mechanical and thermal changes in
nerve during excitation. Physiol. Chem. Phys. Med. NMR, 1988, 20(4), 251 -- 268.
17. Tasaki, I. and Byrne, P. Rapid structural chances in nerve fibers evoked by electric current pulses.
Biochem. Biophys. Res. Commun., 1992, 188(2), 559 -- 564.
|
1705.10805 | 1 | 1705 | 2017-05-30T18:02:35 | Investigations of Auditory Filters Based Excitation Patterns for Assessment of Noise Induced Hearing Loss | [
"physics.bio-ph",
"q-bio.NC"
] | Noise induced hearing loss (NIHL) as one of major avoidable occupational related health issues has been studied for decades. To assess NIHL, the excitation pattern (EP) has been considered as one of mechanisms to estimate movements of basilar membrane (BM) in cochlea. In this study, two auditory filters, dual resonance nonlinear (DRNL) filter and rounded-exponential (ROEX) filter, have been applied to create two EPs, referring as the velocity EP and the loudness EP, respectively. Two noise hazard metrics are also proposed based on the developed EPs to evaluate hazardous levels caused by different types of noise. Moreover, Gaussian noise and pure-tone noise have been simulated to evaluate performances of the developed EPs and noise metrics. The results show that both developed EPs can reflect the responses of BM to different types of noise. For Gaussian noise, there is a frequency shift between the velocity EP and the loudness EP. For pure-tone noise, both EPs can reflect the frequencies of input noise accurately. The results suggest that both EPs can be potentially used for assessment of NIHL. | physics.bio-ph | physics | Investigations of Auditory Filters Based Excitation Patterns for Assessment of
Noise Induced Hearing Loss
Wisam Subhi Al-Dayyeni, Pengfei Sun, Jun Qin
Department of Electrical and Computer Engineering,
Southern Illinois University, Carbondale, IL, USA
Abstract:
Noise induced hearing loss (NIHL) as one of major avoidable occupational related health
issues has been studied for decades. To assess NIHL, the excitation pattern (EP) has been
considered as one of mechanisms to estimate movements of basilar membrane (BM) in cochlea.
In this study, two auditory filters, dual resonance nonlinear (DRNL) filter and rounded-
exponential (ROEX) filter, have been applied to create two EPs, referring as the velocity EP and
the loudness EP, respectively. Two noise hazard metrics are also proposed based on the
developed EPs to evaluate hazardous levels caused by different types of noise. Moreover,
Gaussian noise and pure-tone noise have been simulated to evaluate performances of the
developed EPs and noise metrics. The results show that both developed EPs can reflect the
responses of BM to different types of noise. For Gaussian noise, there is a frequency shift
between the velocity EP and the loudness EP. For pure-tone noise, both EPs can reflect the
frequencies of input noise accurately. The results suggest that both EPs can be potentially used
for assessment of NIHL.
Key words: Noise induced hearing loss; excitation pattern; basilar membrane motion; auditory
filter; noise assessment metrics.
1. Introduction
Noise-induced hearing loss (NIHL) remains as one of the most common health related
problems nowadays as stated by the World Health Organization (WHO). One of the main causes
of the permanent hearing loss is the exposure to excessive noise [1-3]. Approximately 22 million
workers in the United States are exposed to loud noise workplace that is considered as a hazard
level [4]. Hearing loss has a strong impact on the quality of life, causes isolation, impairs social
interactions, and increases the risk of accidents [5].
Intrinsically, NIHL can be partially explained as an auditory fatigue phenomenon, in
which the motions of stretching and squeezing of basilar membrane (BM) could damage the
hearing cells (i.e., outer and inner hair cells) in cochlea [6-8]. The mechanical motions of BM is
considered as one of the major factors that causing NIHL in cochlea [9, 10]. The motions of BM
in response to the noise stimulus as a function of frequency can be stated as an excitation pattern
(EP). Therefore, investigations of the EP are very useful for NIHL research [11].
An EP represents the distribution of movements along BM caused by a sound [12, 13]. In
psychoacoustic, the EP is defined as the output of each auditory filter plotted as a function of the
filter's center frequency (CF) [14]. The EPs are normally calculated and plotted as the gain of
each auditory filter equal to 0 dB at its CF. For example; a tone with a 60 dB sound pressure
level (SPL) and at 1 kHz CF will cause an excitation level equal to 60 dB and at 1 kHz [13, 15,
16].
The auditory models (AMs) of the human peripheral frequency selectivity are the fast
ways to estimate the EPs over the BM partitions in cochlea [17]. Nowadays many AMs have
been developed based on observations of input-output behavior of human auditory system with
reference to psychological or physiological responses [6]. Such AMs include Gammatone filters,
dual-resonance nonlinear (DRNL) filters, dynamic-compressive gammachirp filters, etc.
Hohmann (2002) [18] developed a 4th-order linear Gammatone filter based AM for speech
processing in hearing aids. This linear model can reconstruct acoustical signals in an auditory
system, but it didn't include nonlinear features [18]. Lopez-Poveda and Meddis (2001) [17, 19]
proposed a nonlinear DRNL filter, which successfully simulates the two-tone suppression and
the phase responses in the BM. Furthermore, Irino and Patterson (2006) [20] developed a
gammachirp filterbank with nonlinear and compressive features. The developed gammachirp
filter has a group of linear passive gammachirp filters, and can be useful for the applications on
speech enhancement, speech coding, and hearing aids [20].
Moreover, the AMs can be categorized as mechanical or perceptual model [21]. The
mechanical AMs are designed to estimate mechanical vibrations on BM in cochlea [17], while
the perceptual AMs are developed to mimic the psychoacoustic data [20]. In this study, a DRNL
filter as a typical mechanical AM and a rounded-exponential (ROEX) filter as a typical
perceptual AM have been implemented to investigate EPs on the human BM. As a cascade filter
model, the DRNL filter was developed to simulate the nonlinear mechanical response of BM in
reaction to stapes motion [19]. The output of DRNL filters is the velocity of BM, which can be
described as a velocity EP of BM in cochlea. Such velocity EP intuitively can be used to assess
the auditory fatigue based NIHL [6]. On the other hand, the ROEX filter as a perceptual model
can be used to describe the loudness levels in cochlea. Loudness is one of the most important
parameters for evaluation of the acoustical quality in various applications, from hearing aid
optimizing to automatic music mixing systems [22]. The loudness estimations directly reflect the
characteristics of human auditory system, such as masking adaption, integration along a
perceptual frequency axis, and integration and compression along time axis. In previous studies,
loudness contours based models have been developed for evaluations of the annoyance of
environment noises, including community noise, industrial noise, and transportation noises [23-
25].
In this study, we implement the DRNL filter and the ROEX filter to create two different
EPs, the velocity EP and the loudness EP, respectively. To evaluate the performances of both
EPs, Gaussian noise and pure-tone noise signals with various parameters (e.g., amplitude and
frequency) are simulated. In addition, two noise metrics are proposed based on the velocity EP
and the loudness EP to estimate the hazardous levels caused by different types of noise. The rest
of this paper is organized as follows. Section 2 describes the auditory filters, the proposed noise
metrics, and simulations of noise signals. Section 3 gives experimental results and discussions.
Section 4 concludes the paper and outlines future works.
2. Material and Methods
2.1. External Ear and Middle Ear
The structure and the function of the auditory system of the human have similarities with
other mammalian species. The human ear consists of three parts: external ear, middle ear, and
inner ear. Each part in an auditory system plays a unique role to translate acoustic signals from
environment to inner ear. The sound passes through external ear in a form of pressure vibration.
The pressure vibration changes into mechanical vibration in middle ear. The mechanical energy
transforms into hydrodynamic motion in inner ear, and then the BM activates hair cells through
electrochemical energy [26].
An external ear consists of ear canal, concha, and pinna flange. The transfer functions of
external ear used for the DRNL filter and the ROEX filter are same in this study. As shown in
Fig. 1, the transfer function of external ear is same as it was described in Moore's work [22] and
ANSI-S3-2007 [27].
20
15
10
5
0
-5
)
B
d
(
n
i
a
G
10 2
10 3
10 4
Frequency (Hz)
Fig. 1 - The frequency response of the transfer function of an external ear.
A middle ear consists of tympanic membrane and ossicular chain, which have three
bones. A middle ear plays the role as an impedance-matching device, and it collects and
transmits acoustic power to the inner ear [28, 29]. The transfer functions of middle ear have been
developed to describe the relationships between inputs and outputs of middle ear [29]. In this
study, two different transfer functions of middle ear are applied to the DRNL filter and the
ROEX filter, respectively. Fig. 2a shows the frequency response of the transfer function of
middle ear used for the DRNL filter as described in Meddis's work [19], in which the acoustical
pressure is converted into the stapes velocity, called as the stapes velocity transfer function
(SVTF). As shown in Fig. 2b, the frequency response of the transfer function of middle ear for
the ROEX filter has been used in the procedure of loudness computation in Moore's work [22].
Fig. 2 - The frequency responses of the transfer function of middle ear, which are applied
to (a) the DRNL filter [19] and (b) the ROEX filter [22], respectively.
2.2. DRNL filter
In this study, a DRNL filter is utilized to obtain the BM movements in human cochlea
[19]. The DRNL filter simulates the velocity of BM as a response to the stapes velocity in middle
ear. As shown in Fig. 3, the input of the DRNL filter is the linear stapes velocity. Each individual
site is represented as a tuned system with two parallel independent paths, one linear (left) and
one nonlinear (right). The linear path consists of a gain /attenuation factor, a bandpass function,
and a low pass function in a cascade. The nonlinear path is a cascade combination of the 1st
bandpass function, a compression function, the 2nd bandpass function, and a low pass function.
The output of DRNL filter is the sum of the outputs of the linear and nonlinear paths, and is the
BM velocity at a particular location along the cochlear partition.
In both paths, each of three bandpass functions consists of a cascade of two or three the
1st order gammatone filters [30] with a unit gain at the center frequency (CF). Two low-pass
functions are same and consist of a cascade of four 2nd order Butterworth low pass filters.
Moreover, the compression function in the nonlinear path was defined based on the
animal data, and it can be described as
𝑦[𝑡]=SIGN (𝑥[𝑡])×MIN (𝑎𝑥[𝑡],𝑏𝑥[𝑡])𝑐
where 𝑥[𝑡] represents the output from the first bandpass function in the nonlinear path. 𝑦[𝑡] is
the output of the compression function. 𝑎, 𝑏, and 𝑐 are models parameters as summarized in
Table 1.
(1)
Fig. 3 - Schematic diagram of the DRNL filter [19], in which the velocities of stapes in
middle ear are passed through two parallel branches to obtain the velocities of BM.
Table 1 summarizes the parameters of the DRNL filter for human used to implement the
DRNL filter in this study. The velocity EP is the distribution of BM velocity, which can be
obtained as the outputs of the DRNL filter.
Table 1 - DRNL filter parameters used to simulate the human inner ear [17].
0.25kHz
0.5kHz
1kHz
2kHz
4kHz
8kHz
2
4
235
115
235
1400
3
3
250
84
250
2124
0.45
0.25
2
4
460
150
460
800
3
3
500
103
500
4609
0.28
0.25
2
4
945
240
945
520
3
3
1000
175
1000
4598
0.13
0.25
2
4
1895
390
1895
400
3
3
2000
300
2000
9244
0.078
0.25
2
4
3900
620
3900
270
3
3
4000
560
4000
30274
0.06
0.25
2
4
7450
1550
7450
250
3
3
8000
1100
8000
76354
0.035
0.25
Simulated
preparation
Linear
GT cascade
LP cascade
CFlin
BWlin
LPlin
Gain, g
Nonlinear
GT cascade
LP cascade
CFlin
BWlin
LPnl
Gain, a
Gain, b
Exponent, c
2.3. ROEX filter
The ROEX filter was originally derived from psychophysical data [31]. It is a descriptive
model, which describes the shape of magnitude transfer function of an auditory filter [32]. The
ROEX filter formula can be defined by [27]:
𝑊(𝑔)= (1+𝑝𝑔)𝑒𝑥𝑝(−𝑝𝑔)
where 𝑔 is the normalized deviation from the center frequency (CF) divided by the CF, and 𝑝 is
an adjustable parameter which determines the slope and the bandwidth of the filter.
(2)
In this study, the ROEX filter is implemented according to ANSI 3.4-2005 [33]. To
calculate the input level at each equivalent rectangular bandwidth (ERB), 𝑝 in Eq. (3) is set to be
4𝑓/𝐸𝑅𝐵. The ERB is a psychoacoustic measurement of the width of the auditory filter in each
where 𝑓𝑐 is the CF, which are in the range of 50 Hz – 15 kHz in this study. The ERB level
𝐸𝑅𝐵 = 24.673(0.004368𝑓𝑐+1)
obtained according to the input level, which is used to determine the ROEX filter shape. The
location along the cochlea, and it can be defined as:
(3)
energy in each ERB can be obtained by:
𝐸𝑖=∑𝑊�𝑔𝑖,𝑗�𝑃𝑗2
𝑃02
𝐸0
(4)
according to the values of the excitation threshold
where 𝑊 represents local ROEX filter in the ith ERB. 𝑃𝑗2 refers to the power in the jth frequency
band. 𝐸0 is the reference energy at 1 kHz CF and 0 dB SPL, and 𝑃0 is the reference pressure
referring to 2×10−5 Pa. For the selected frequencies, Ei will be transformed to loudness levels
where 𝐸 is the energy, and G is the low level gain. 𝐶 and 𝛼 are two constants, where 𝐶
=0.046871, and α is related to the 𝐺 value. 𝐸𝑇𝐻𝑅𝑄 refers to lower threshold of human
𝑁=𝐶��𝐺×𝐸 + 2𝐸𝑇𝐻𝑅𝑄 �α – �2𝐸𝑇𝐻𝑅𝑄�α�
(5)
perception.
Fig. 4 shows various frequency gains of ROEX filter at different ERB levels. The ROEX
filter is a dynamic filter, which has different frequency gains when the levels of ERB change. As
the ERB level increases, the slope of the left side of ROEX filter becomes flat. In general, when
the sound pressure level (SPL) increases, there will be more energy pass through the ROEX
filter. From this perspective, ROEX filter is consistent with the loudness contours. When SPLs
increase, loudness contours become flat [14].
Fig. 4 - The various frequency gains of ROEX filter centered at 0.5, 1, 2, 4, and 8 kHz at
different ERB levels from 20 dB to 100 dB with 10 dB interval.
2.4. EP Based Noise Metrics
Previous studies have demonstrated that the EPs of BM are highly correlated with NIHL
in human cochlea [6, 11, 34]. To investigate hearing loss, two EP based metrics are proposed to
assess the potential hazardous levels (HLs) caused by different types of noise. Since the EP
represents the temporal responses of the organ of Corti in cochlea, one can integrate the local
responses and obtain the cumulative HLs. Therefore, two proposed noise metrics, 𝐻𝐿𝑖𝐷 and
𝐻𝐿𝑖𝑅, can be defined as
(6)
𝐻𝐿𝑖𝐷=10log10� 𝑉(𝑖,𝑡)2 𝑉𝑜2
�
𝐻𝐿𝑖𝑅=10log10� 𝑁(𝑖,𝑡)2 𝑁𝑜2
�
𝑡=𝑛𝑡=1
𝑡=𝑛𝑡=1
(7)
where 𝐻𝐿𝑖𝐷 represents the hazard level index based on the velocity EP, and 𝑉(𝑖,𝑡) refers to the
BM velocity at the 𝑖th ERB of BM at time 𝑡. 𝑉0 represents the BM velocity located at the ERB at
CF equal to 1 kHz. Moreover, 𝐻𝐿𝑖𝑅 represents the hazard level index based on the loudness EP,
and 𝑁(𝑖,𝑡) refers to the loudness level at the ith ERB of BM at a time t. 𝑁0 is the loudness level
at the ERB at CF equal to 1 kHz. By Eq. (7) and Eq. (8), the developed EPs have been
successfully translated to the amount of HLs, which can be potentially used for the assessment of
NIHL.
Moreover, total hazard level (THL) can be defined as summation of HLs:
where 𝑇𝐻𝐿 𝐷 and 𝑇𝐻𝐿 𝑅 represent THLs based on the velocity EP and the loudness EP,
respectively.
2.5. Simulation of Noise Signals
In this study, two different types of noise signals (i.e., Gaussian noise and pure-tone
noise) have been simulated to evaluate the performances of two developed EPs. The Gaussian
noise signals are simulated using the "randn" function in MATLAB, in which the probability
distribution function of the Gaussian noise is given by [35]:
𝑃(𝑡)= 1𝜎√2𝜋𝑒𝑥𝑝−(𝑡−𝜇)2
2𝜎2
where 𝜇 is the mean, and σ is the standard deviation. 𝜇 is equal to zero in this study.
(10)
𝑇𝐻𝐿 𝐷= ∑𝐻𝐿𝑖𝐷
𝑖
𝑇𝐻𝐿 𝑅= ∑𝐻𝐿𝑖𝑅
𝑖
(8)
(9)
The pure-tone noise signals are simulated by:
𝑦(𝑡)=𝐴cos2𝜋𝑓𝑡
(11)
where 𝐴 is the amplitude of the signal, and 𝑓 is the frequency.
3. Results and Discussion
3.1. Time-Frequency (T-F) Representations of Two EPs
In this section, two simulated noise signals (i.e., Gaussian noise and pure-tone noise) are
fed into both velocity EP and loudness EP models. The outputs of two EP models are the BM
velocity and the loudness level 𝑁(𝑖,𝑡) at the 𝑖th ERB of BM at time 𝑡, respectively. Both EP can
be represented in the joint time and frequency (T-F) domain. Figs. 5a and 5b show the T-F
representations of the velocity EP and the loudness EP, responding to a simulated Gaussian noise
at 100 dB SPL, respectively. Figs. 5c and 5d show the T-F representations of the velocity EP and
the loudness EP, responding to a pure-tone noise with 100 dB SPL and 1 kHz frequency,
respectively. The results show that both EPs can reflect amplitudes and transitions of noise
signals. The velocity EP as a mechanical model can represent both positive and negative
vibrations of BM in cochlea, which reflects more realistic representations of the stretching and
squeezing on the hair cells in cochlea. In the other hand, the loudness EP as a perceptual model
only represents the positive amount of the loudness as a response to the noise signal. The
loudness EP doesn't directly reflect the BM vibrations in cochlea.
Fig. 5 - The T-F representations of (a) the velocity EP and (b) the loudness EP responding
to a Gaussian noise at 100 dB SPL, and (c) the velocity EP and (d) the loudness EP with
respect to a pure-tone noise at 100 dB SPL and 1 kHz.
Moreover, along the time axis, the velocity EP presents higher temporal resolution than
the loudness EP for both Gaussian and pure-tone noise. It indicates that the temporal resolution
of the DRNL filter is better than that of the ROEX filter. Along the frequency axis, for the
Gaussian noise case, the peak frequency of the velocity EP is around 2 kHz and is lower than the
corresponding value of the loudness EP (around 4 kHz). For the pure-tone noise, both EPs
present the peak frequencies at 1 kHz, which reflects the frequency of the input pure-tone noise.
However, the velocity EP shows vibrations around 1 kHz since it reflects the BM motion while
the loudness EP shows only one pulse since it is a perceptual model that reflects the amount of
psychoacoustic data.
3.2. T-F Representations of Two EPs for Pure-tone Noise
Fig. 6 shows the T-F representations of the velocity EP and the loudness EP produced by
the pure-tone noise signals with 100 dB SPL and various frequencies (i.e., 1, 2, 4, and 6 kHz).
For the velocity EP (as shown in the left figures of Fig. 6), the amplitudes have both positive and
negative values and the peak amplitudes appear around the frequencies of pure-tone noise
signals. It also can be found that the peak amplitudes of the velocity EP are decreasing with the
frequency large than 2 kHz. On the other hand, the loudness EP (as shown in the right figures of
Fig. 6) presents only positive amplitudes, and the peak amplitudes match the frequencies of
stimulating pure-tone noise signals. The peak amplitudes of the loudness EP increase first and
then decrease with the frequency increasing, and the maximum peak amplitude appears at 4 kHz.
Fig. 6 – The T-F distributions of two developed EPs obtained by simulated pure-tone noise
signals at 100 dB SPL with frequencies at 1, 2, 4, and 6 kHz, respectively.
3.3.Hazardous Level Evaluation
3.3.1. Frequency Distributions of HLs for Gaussian Noise
According to Eqs. (6) and (7), the performance two EPs are evaluated using two proposed
metrics, 𝐻𝐿𝑖𝐷and 𝐻𝐿𝑖𝑅 , which are used to depict HL at ith ERB on BM. Fig. 7 shows the
frequency distributions of normalized HLs generated by the simulated Gaussian noise signals at
SPL = 90 to 120 dB with 10 dB interval. For both velocity EP and loudness EP, the HLs increase
with SPL increasing. Overall, the loudness EP shows broader frequency response compared with
the velocity EP. The results also show that there is a frequency shift between the two EPs. The
peak HLs of the velocity EP are around 2 kHz while the peak HLs of the loudness EP are around
4 kHz. Since the BM motions are associated with hearing loss in cochlea, the peak frequency
shift between two EPs indicates that the maximum hearing loss predicted by these two EPs may
occur at different partitions of BM.
Fig. 7 - The frequency distributions of normalized HLs based on (a) the velocity EP and (b)
the loudness EP generated by simulated Gaussian noise signals at SPL = 90 to 120 dB with
10 dB interval.
3.3.2. Frequency Distributions of HLs for Pure-tone Noise
Fig. 8 shows the normalized HLs generated by simulated pure-tone noise signals at 1 kHz
fixed frequency and SPL from 90 to 120 dB with 10 dB interval. Both velocity EP and loudness
EP show the peak frequency responses at 1 kHz, which is same as the frequency of the input
pure-tone noise signals. It also can be found that the HLs are increasing with SPL levels
increasing in both EPs. As shown in Fig. 8a, the HLs of the velocity EP gradually increase when
the frequency is smaller than 1 kHz, and then gradually decrease after the frequency is greater
than 1 kHz. Comparatively, as shown in Fig. 8b, the HLs of the loudness EP show different
frequency responses than the velocity EP. The HLs of the loudness EP almost equal to zero when
frequency smaller than 500 Hz, and then rapidly increase with frequency increasing to 1 kHz,
and finally gradually decrease with frequency further increasing. This is because the loudness EP
is based on the ROEX filter, which is derived from psychophysical data. Therefore, the loudness
EP may not reflect the real motion of BM in cochlea.
Fig 8 - The frequency distributions of normalized hazardous levels based on (a) the velocity
EP and (b) the loudness EP obtained pure-tone noise at 1 kHz and SPL = 90 to 120 dB with
10 dB interval.
Moreover, Fig. 9 shows the normalized HLs generated by the simulated pure-tone noise
signals at different frequencies (0.5, 1, 2, 4, and 6 kHz) and SPL = 100 dB. Both velocity EP and
loudness EP can reflect the corresponding frequencies of input pure-tone noise signals.
Specifically, the peak HLs of the velocity EP (as shown in Fig. 9a) is reducing after frequency
larger than 2 kHz, while the peak HLs of the loudness EP slightly reduce when frequency higher
than 4 kHz. The results in Fig. 9 confirm the peak frequency shift between two EPs generated by
Gaussian noise in Fig. 7. In the velocity EP, the maximum velocity occurs around 2 kHz, while
in the loudness EP, the maximum loudness appears around 4 kHz.
Fig 9 - The frequency distributions of normalized hazardous levels based on (a) the velocity
EP and (b) the loudness EP obtained pure-tone noise signals at various frequencies (0.5, 1,
2, 4, and 6 kHz) with fixed SPL = 100 dB.
3.3.3. Total Hazardous Levels for Gaussian Noise
According to Eq. (8) and Eq. (9), total hazardous levels, 𝑇𝐻𝐿 𝐷 and 𝑇𝐻𝐿 𝑅 , can be
calculated based on the velocity EP and the loudness EP, respectively. THLs can be used to
assess hazardous of high-level noise, and potentially can be used to investigate NIHL. Fig. 10
shows the evaluation of the normalized THLs for the Gaussian noise at SPL from 70 to 120 dB.
The result shows that THLs of both EPs are increasing with SPL increasing. The increase rate of
the velocity EP is faster than that of the loudness EP. Compared with the loudness EP, the
velocity EP shows lower THLs at SPL < 100 dB, but demonstrates higher THLs when SPL >
100 dB.
Fig 10 - The normalized THLs for the Gaussian noise at SPL from 70 to 120 dB for the
velocity EP and the loudness EP.
3.3.4. THLs for the Pure-tone Noise
Fig. 11a shows the normalized THL of both EPs produced by the simulated pure-tone
noise signals with increasing SPL from 70 to 120 dB and fixed frequency at 1 kHz. The THLs of
both EPs are increasing with SPL increasing. Specifically, the velocity EP increases faster than
the loudness EP. The result indicates that the velocity EP is more sensitive with SPL increasing
than the loudness EP. It also can be found that the THLs of the velocity EP are constantly higher
than the corresponding values of the loudness.
Moreover, Fig. 11b shows the normalized THLs of both EPs generated by the simulated
pure-tone noise signals at SPL = 100 dB and frequency from 0.5 to 8 kHz. For both EPs, the
THLs slightly increase first and then decrease with frequency increasing. The peak THL of the
velocity EP is at 2 kHz, while the THL of the loudness EP peaks at 4 kHz. This result is
consistent with the previous results in Figs 6, 7 and 9. In addition, the velocity EP shows a fast
degradation
of THL when
the
frequency
increase more
than
2
kHz.
Fig 11 - The normalized THLs for the pure-tone noise: (a) at 1 kHz and SPL from70 to120
dB, and (b) at fixed SPL = 100 dB and frequencies from 0.5 to 8 kHz for the velocity EP
and the loudness EP.
4. Conclusions
In this study, two auditory filters, the DRNL filter and the ROEX filter, have been
applied to develop the velocity EP and the loudness EP, respectively. Two different types of
noise (i.e., Gaussian noise and pure-tone noise) have been simulated to evaluate two developed
EPs. For the Gaussian noise, the results show that the maximum velocity obtained by the DRNL
filter occurs around 2 kHz, while the peak loudness obtained by the ROEX filter is about 4 kHz.
For the pure-tone noise, both EPs can accurately reflect the frequencies of the input noise
signals. Moreover, to evaluate the effectiveness of two EPs for prediction of NIHL, we proposed
two noise metrics, 𝐻𝐿𝐷 and 𝐻𝐿𝑅, based on the velocity EP and the loudness EP, respectively.
The results show that both EPs can be potentially used as noise hazardous level index for
assessment of NIHL. The velocity EP based metric demonstrates higher sensitivity than the
loudness EP based metric. However, because the current study is only based on theoretical
analysis and simulated noise signals, it may be limited to evaluate the performance of two
auditory filters. In our future work, we will utilize experimental animal and human noise
exposure data to evaluate the developed velocity EP and loudness EP for assessment of NIHL.
References
journal of
[1] Rabinowitz PM. Noise-induced hearing loss. American family physician. 2000;61:2759-60.
[2] Qin J, Jiang Y, Mahdi A. Recent Developments on Noise Induced Hearing Loss for Military and
Industrial Applications. Biosens J. 2014;3:e101.
[3] Wu Q, Qin J. Effects of key parameters of impulse noise on prediction of the auditory hazard using
AHAAH model. International journal of computational biology and drug design. 2013;6:210-20.
[4] Tak S, Davis RR, Calvert GM. Exposure to hazardous workplace noise and use of hearing protection
devices among US workers-NHANES, 1999–2004. American
industrial medicine.
2009;52:358-71.
[5] Kirchner DB, Evenson E, Dobie RA, Rabinowitz P, Crawford J, Kopke R, et al. Occupational noise-
induced hearing loss: ACOEM Task Force on occupational hearing loss. Journal of Occupational and
Environmental Medicine. 2012;54:106-8.
[6] Sun P, Qin J, Campbell K. Fatigue Modeling via Mammalian Auditory System for Prediction of Noise
Induced Hearing Loss. Computational and mathematical methods in medicine. 2015;2015.
[7] Sun P, Fox D, Campbell K, Qin J. Auditory fatigue model applications to predict noise induced hearing
loss in human and chinchilla. Applied Acoustics. 2017;119:57-65.
[8] Sun P, Qin J. Auditory fatigue models for prediction of gradually developed noise induced hearing
loss. 2016 IEEE-EMBS International Conference on Biomedical and Health Informatics (BHI)2016. p. 384-
7.
[9] Calford M, Rajan R, Irvine D. Rapid changes in the frequency tuning of neurons in cat auditory cortex
resulting from pure-tone-induced temporary threshold shift. Neuroscience. 1993;55:953-64.
[10] Ohlemiller KK. Contributions of mouse models to understanding of age-and noise-related hearing
loss. Brain research. 2006;1091:89-102.
[11] Sun P, Qin J. Excitation patterns of two auditory models applied for noise induced hearing loss
assessment. Biomedical and Health Informatics (BHI), 2016 IEEE-EMBS International Conference on:
IEEE; 2016. p. 398-401.
[12] Fletcher H. Auditory patterns. Reviews of modern physics. 1940;12:47.
[13] Chen Z, Hu G, Glasberg BR, Moore BC. A new method of calculating auditory excitation patterns and
loudness for steady sounds. Hearing research. 2011;282:204-15.
[14] Moore BC, Glasberg BR. Suggested formulae for calculating auditory‐filter bandwidths and
excitation patterns. The Journal of the Acoustical Society of America. 1983;74:750-3.
[15] Zwicker E. Masking and psychological excitation as consequences of the ear's frequency analysis.
Frequency analysis and periodicity detection in hearing. 1970:376-94.
[16] Glasberg BR, Moore BC. Derivation of auditory filter shapes from notched-noise data. Hearing
research. 1990;47:103-38.
[17] Lopez-Poveda EA, Meddis R. A human nonlinear cochlear filterbank. The Journal of the Acoustical
Society of America. 2001;110:3107-18.
[18] Hohmann V. Frequency analysis and synthesis using a Gammatone filterbank. Acta Acustica united
with Acustica. 2002;88:433-42.
[19] Meddis R, O'Mard LP, Lopez-Poveda EA. A computational algorithm for computing nonlinear
auditory frequency selectivity. The Journal of the Acoustical Society of America. 2001;109:2852-61.
[20] Irino T, Patterson RD. A dynamic compressive gammachirp auditory filterbank. Audio, Speech, and
Language Processing, IEEE Transactions on. 2006;14:2222-32.
[21] Saremi A, Beutelmann R, Dietz M, Ashida G, Kretzberg J, Verhulst S. A comparative study of seven
human cochlear filter models. The Journal of the Acoustical Society of America. 2016;140:1618-34.
[22] Moore BC, Glasberg BR, Baer T. A model for the prediction of thresholds, loudness, and partial
loudness. Journal of the Audio Engineering Society. 1997;45:224-40.
[23] Schomer PD, Suzuki Y, Saito F. Evaluation of loudness-level weightings for assessing the annoyance
of environmental noise. The Journal of the Acoustical Society of America. 2001;110:2390-7.
[24] Sun P, Qin J, Qiu W. Development and validation of a new adaptive weighting for auditory risk
assessment of complex noise. Applied Acoustics. 2016;103:30-6.
[25] Qin J, Sun P, Walker J. Measurement of field complex noise using a novel acoustic detection system.
AUTOTESTCON, 2014 IEEE: IEEE; 2014. p. 177-82.
[26] Johnson K. Acoustic and auditory phonetics. Phonetica. 2004;61:56-8.
[27] ANSI S. 4: Procedure for the computation of loudness of steady sounds. ANSI S3. 2007:4-2007.
[28] Rosowski JJ. Outer and middle ears. Comparative Hearing: Mammals: Springer; 1994. p. 172-247.
[29] Slama MC, Ravicz ME, Rosowski JJ. Middle ear function and cochlear input impedance in chinchilla.
The Journal of the Acoustical Society of America. 2010;127:1397-410.
[30] Hartmann WM. Signals, sound, and sensation: Springer Science & Business Media; 1997.
[31] Patterson RD, Nimmo‐Smith I, Weber DL, Milroy R. The deterioration of hearing with age:
Frequency selectivity, the critical ratio, the audiogram, and speech threshold. The Journal of the
Acoustical Society of America. 1982;72:1788-803.
[32] Patterson RD. Auditory filter shapes derived with noise stimuli. The Journal of the Acoustical Society
of America. 1976;59:640-54.
[33] Acoustical Society of A, Acoustical Society of A, Standards S, American National Standards I,
Accredited Standards Committee S B. American National Standard : procedure for the computation of
loudness of steady sounds. Melville, N.Y.: Standards Secretariat, Acoustical Society of America; 2005.
[34] Crane H. Mechanical impact: a model for auditory excitation and fatigue. The Journal of the
Acoustical Society of America. 1966;40:1147-59.
[35] Hussain ZM, Sadik AZ, O'Shea P. Digital signal processing: an introduction with MATLAB and
applications: Springer Science & Business Media; 2011.
|
1211.0700 | 1 | 1211 | 2012-11-04T16:17:49 | A transient solution for vesicle electrodeformation and relaxation | [
"physics.bio-ph"
] | A transient analysis for vesicle deformation under DC electric fields is developed. The theory extends from a droplet model, with the additional consideration of a lipid membrane separating two fluids of arbitrary properties. For the latter, both a membrane-charging and a membrane-mechanical model are supplied. The vesicle is assumed to remain spheroidal in shape for all times. The main result is an ODE governing the evolution of the vesicle aspect ratio. The effects of initial membrane tension and pulse length are examined. The model prediction is extensively compared with experimental data, and is shown to accurately capture the system behavior in the regime of no or weak electroporation. More importantly, the comparison reveals that vesicle relaxation obeys a universal behavior regardless of the means of deformation. The process is governed by a single timescale that is a function of the vesicle initial radius, the fluid viscosity, and the initial membrane tension. This universal scaling law can be used to calculate membrane properties from experimental data. | physics.bio-ph | physics |
Transient solution for vesicle electrodeformation and relaxation
A transient solution for vesicle electrodeformation and relaxation
Jia Zhang,1 Jeffery D. Zahn,2 Wenchang Tan,3 and Hao Lin1, a)
1Department of Mechanical and Aerospace Engineering, Rutgers, The State University of New Jersey, Piscataway,
NJ 08854, USA
2Department of Biomedical Engineering, Rutgers, The State University of New Jersey, Piscataway, NJ 08854,
USA
3State Key Laboratory for Turbulence and Complex Systems, Department of Mechanics and Engineering Science,
College of Engineering, Peking University, Beijing 100871, China
(Dated: 1 November 2018)
A transient analysis for vesicle deformation under DC electric fields is developed. The theory extends from
a droplet model, with the additional consideration of a lipid membrane separating two fluids of arbitrary
properties. For the latter, both a membrane-charging and a membrane-mechanical model are supplied. The
vesicle is assumed to remain spheroidal in shape for all times. The main result is an ODE governing the
evolution of the vesicle aspect ratio. The effects of initial membrane tension and pulse length are examined.
The model prediction is extensively compared with experimental data, and is shown to accurately capture the
system behavior in the regime of no or weak electroporation. More importantly, the comparison reveals that
vesicle relaxation obeys a universal behavior regardless of the means of deformation. The process is governed
by a single timescale that is a function of the vesicle initial radius, the fluid viscosity, and the initial membrane
tension. This universal scaling law can be used to calculate membrane properties from experimental data.
I.
INTRODUCTION
Vesicles are widely used as a model system for biolog-
ical cells due to their simplicity and controllability. The
deformation of the lipid membrane, in particular under
an applied electric field (electrodeformation), is often ex-
plored to probe membrane properties1,2 and to detect
pathological changes in cells.3
In the past decade, vesicle electrodeformation has be-
come a significant subject of study, and earlier work
can be divided into two categories.
In the first cate-
gory, an alternating-current (AC) field is applied, which
often induces stationary and small deformations.1,2,4,5
Correspondingly, an electrohydrodynamic theory in the
small-deformation limit was developed to interpret the
data trends.6 In the second, under direct-current (DC)
electric fields, vesicles usually exhibit large and tran-
sient deformations due to the large field strengths com-
monly applied.7 -- 10 Recently, using high-resolution, high-
speed optical imaging Riske and Dimova8 acquired a
large amount of data capturing the complex deformation-
relaxation behavior of the vesicles. Although some qual-
itative and scaling arguments were presented,4 the data
was not fully interpreted due to the absence of a pre-
dictive model. Meanwhile, one of us (HL) experimen-
tally examined vesicles in the large-deformation regime
with aspect ratios reaching ten.10 A large-deformation
theory was also presented, which provided quantitative
agreement with the data therein. However, the model
was semi-empirical in that the hydrodynamic problem
was not rigorously treated, but followed an empirical ap-
proach by Hyuga and co-authors.11,12 In general, a rig-
orous and transient analysis needs to be developed to
a)Email address for correspondence: [email protected]
understand the complex deformation-relaxation behav-
ior, and to provide insights on the underlying physical
processes.
In this work, we develop a transient analysis for vesi-
cle electrodeformation. The theory is derived by extend-
ing our previous work on a droplet model,13 with the
additional consideration of a lipid membrane separating
two fluids of arbitrary properties. For the latter, both a
membrane-charging and a membrane-mechanical model
are supplied. Similar to the droplet model, the main re-
sult is also an ordinary differential equation (ODE) gov-
erning the evolution of the vesicle aspect ratio. The ef-
fects of initial membrane tension and pulse length are
examined. The model prediction is extensively com-
pared with experimental data from Riske and Dimova8
and Sadik et al.,10 and is shown to accurately capture
the system behavior in the regime of no or weak elec-
troporation. More importantly, the comparison reveals
that vesicle relaxation obeys a universal behavior, and
is governed by a single timescale that is a function of
the vesicle initial radius, the fluid viscosity, and the ini-
tial membrane tension. This behavior is regardless of the
means of deformation, either via AC/DC electric field, or
via mechanical stretching. This universal scaling law is a
main contribution of the current work, and can be used to
calculate membrane properties from experimental data.
II. THEORY
The problem configuration is shown in Fig. 1. Under
the influence of an applied electric field, charges of op-
posite signs are allowed to accumulate on the two sides
of the membrane, which induces vesicle deformation and
electrohydrodynamic flows both inside and outside the
vesicle. We assume that the vesicle remains spheroidal
in shape throughout the process. All notations, as well
Transient solution for vesicle electrodeformation and relaxation
2
as the prolate spheroidal coordinate system follow those
from Zhang et al..13 The surface of the prolate spheriod
is conveniently given as
ξ = ξ0 ≡
a
c
.
(1)
Here c = √a2 − b2 is chosen to be the semi-focal length
of the spheroidal vesicle, and a and b are the major and
minor semi-axis, respectively. For the derivation below,
we further assume that the volume of the vesicle is con-
served. We subsequently obtain
a = r0(1 − ξ−2
0 )−
1
3 ,
b = r0(1 − ξ−2
0 )
1
6 .
(2)
Therefore, the vesicle geometry is completely character-
ized by a single parameter, ξ0, which evolves in time along
with deformation. The critical idea of the current anal-
ysis is to express all variables, e.g., the electric potential
and the stream function in terms of ξ0. In what follows,
we introduce both an electrical and a mechanical model
for the membrane. An ODE for ξ0 is obtained by apply-
ing the stress matching and kinematic conditions.
A. The electrical problem
The electric potentials both inside and outside the vesi-
cle are described by the Laplace equations:
∇2φi = ∇2φe = 0.
(3)
However, at the membrane the matching conditions are
modified:
σe
hξ
∂φe
∂ξ
=
σi
hξ
∂φi
∂ξ
=Cm
∂ c
hξ
(φe − φi)
∂t
+
Gmc
hξ
(φe − φi),
at ξ = ξ0.(4)
Here Cm and Gm denote the membrane capacitance and
conductance, respectively. hξ is a metric coefficient of the
prolate spheroidal coordinate system. This membrane-
charging model is commonly adopted by many previ-
ous research.14 -- 18 The displacement currents from the
electrolytes are not included, which approximation is
valid when the Maxwell-Wagner timescale, TMW = (ǫi +
2ǫe)/(σi + 2σe), and the charge relaxation timescale,
Tcr = ǫ/σ, are small when compared with the membrane-
charging time, Tch = r0Cm(1/σi + 1/2σe), and the de-
formation time, Td = µe/ǫeE2
0 . However, the last two
times are in general comparable with each other. The
first term on the RHS of Eq. (4) represents capacitive
charging of the membrane, which includes the effect of
membrane deformation. However, the contribution from
this effect is usually small, and is neglected in the current
analysis for simplicity. Equation (4) can be consequently
reduced to
∂φe
∂ξ
∂(φe − φi)
∂φi
∂ξ
∂t
=
σe
hξ
=
σi
hξ
Cmc
hξ
Gmc
hξ
+
(φe − φi),
at ξ = ξ0.(5)
FIG. 1. (a) A schematic of the problem configuration. The
original radius of the vesicle is r0. The conductivity is denoted
by σ, the permittivity is denoted by ǫ, the viscosity is denoted
by µ, and the subscripts i and e denote intravesicular and ex-
travesicular, respectively. The strength of the applied electric
field is E0. (b) The prolate spheroidal coordinate system.
Equation (5) can be further simplified by considering dif-
ferent stages of charging. In the first stage, the trans-
membrane potential (TMP), Vm ≡ (φi − φe)ξ=ξ0 , grows
continuously in magnitude, but the membrane is not per-
meabilized. Under this condition, Gm is near zero, and
Eq. (5) becomes
σe
hξ
∂φe
∂ξ
=
σi
hξ
∂φi
∂ξ
=
Cmc
hξ
∂(φe − φi)
∂t
,
at ξ = ξ0. (6)
In the second stage, the maximum TMP reaches the
critical threshold, Vc, for electroporation to occur.19 -- 25
The membrane becomes permeable to ions, and Gm in-
creases significantly to limit further growth of the TMP.
In general, the exact values of Vm and Gm depend on
the detailed electroporation conditions and variables such
as pore density and pore area.18 The solution usually
requires a complex numerical simulation which is be-
yond the scope of the theoretical analysis pursued in
this paper. However, a comprehensive model study by
Li and Lin18 showed that the maximum TMP remained
at the critical level in the presence of the pulse post-
permeabilization. In this work, we adopt an approximate
Transient solution for vesicle electrodeformation and relaxation
3
model for this stage. We assume that once the maximum
value of Vm reaches Vc, it no longer grows and "freezes"
in time. In addition, the membrane is completely perme-
abilized, and Eq. (5) is replaced by
σe
hξ
∂φe
∂ξ
=
σi
hξ
∂φi
∂ξ
,
Vm = Vc,
at ξ = ξ0.
(7)
Note that electroporation only occurs for sufficiently
strong electric fields, and Eq. (7) is not needed for some
of the cases studied below where Vc is never reached. Far
away from the vesicle surface, the electric field is uniform
− ∇φe = E0z,
at ξ → ∞.
(8)
We also require that φi remains finite at ξ = 1. For initial
condition, we solve Eqs. (3) and (6) with Vm = 0.
The general solution of the electric potentials for both
the exterior and interior of the vesicle can be obtained
following a similar procedure outlined in Zhang et al.:13
φe = E0r0 [−λξ + αQ1(ξ)] η,
φi = E0r0βξη.
(9)
(10)
Here, Q1(ξ) is a 1st-degree Legendre polynomial of the
second kind. λ ≡ c/r0 is the dimensionless semi-focal
length. The coefficients α and β are again obtained by
applying the matching conditions. In the absence of elec-
troporation, they are given as
α =
β + σrλ
Q′
1(ξ0)σr
,
1(ξ0)σr − ξ0(cid:21) dβ
(cid:20) Q1(ξ0)
Q′
dτ −" Q1(ξ0)Q
−"(cid:18)ξ0 −
′′
1 (ξ0) − Q
1 (ξ0)(1 − σr)
′2
Q′2
1 (ξ0)σr
dξ0
dτ
+
Q1(ξ0)
Q′
1(ξ0)(cid:19) dλ
dξ0
λQ
+
′′
1 (ξ0)Q1(ξ0)
Q′2
1 (ξ0)
τ2
τ1λ# β
# dξ0
dτ
= 0,
α(0) =
λξ0(σr − 1)
Q′
1(ξ0)ξ0σr − Q1(ξ0)
,
β(0) =h−λ + α(0)Q
′
1(ξ0)i σr.
(11)
(12)
(13)
Here σr ≡ σe/σi is the conductivity ratio. τ1 ≡ r0Cm/σi
is a membrane-charging time. τ2 ≡ r0µe/Γ0 is a charac-
teristic flow timescale. Γ0 is the initial membrane ten-
sion introduced below. The dimensionless time τ defined
as τ ≡ t/τ2 has been used. Note that the definition of
these times slightly deviates from those used in Zhang et
al.13 due to the difference between droplet and vesicle.
However, τ2 remains formally the same by replacing γ in
Zhang et al.13 with Γ0.
After the maximum value of Vm reaches the critical
threshold, electroporation occurs. α and β are calculated
by Eq. (7) which yields
α = −Vc/(E0r0) − λξ0(σr − 1)
Q1(ξ0) − Q′
1(ξ0)ξ0σr
, β =h−λ + αQ
′
1(ξ0)i σr.
(14)
The expressions for the normal and tangential electro-
static stresses are found in Zhang et al.13 and not re-
peated here.
B. The hydrodynamic problem
In the regime of low-Reynolds-number flow, the gov-
erning equation for the hydrodynamic problem can be
rewritten in terms of the stream function, ψ, as
E4ψ = 0.
(15)
Here, the expression for the operator E2 can be found in
Dubash and Mestel26 and Bentenitis and Krause.27 The
stream function is related to the velocity components as
u = −
1
hξhθ
∂ψ
∂ξ
,
v =
1
hηhθ
∂ψ
∂η
.
(16)
hη and hθ are metric coefficients of the prolate spheroidal
coordinate system. At the membrane, u and v represent
the tangential and normal velocities, respectively, and
they are required to be continuous
ue = ui,
ve = vi,
at ξ = ξ0.
(17)
In addition, we prescribe a kinematic condition relating
the membrane displacement to the normal velocity,
v(ξ = ξ0, η) =
r0(cid:0)1 − ξ−2
0 (cid:1)−5/6
3ξ2
0
dξ0
dt
.
(18)
(cid:0)1 − 3η2(cid:1)
pξ2
0 − η2
At the membrane, the stress matching condition is given
as:
τ · n = f mem.
(19)
Transient solution for vesicle electrodeformation and relaxation
4
Γ0 = 10−9 N/m
Γ0 = 10−7 N/m
Γ0 = 10−5 N/m
Γ0 = 10−3 N/m
0.25
0.6
0.2
0.5
∆
0.15
0.4
0.3
0
0.1
0.05
0.05
0.1
0
10−9 10−8 10−7 10−6 10−5 10−4 10−3 10−2 10−1
Γh (N/m)
FIG. 2. The relative increase of the apparent area, ∆, as a
function of membrane tension, Γh, for different values of ini-
tial membrane tension, Γ0. The inset shows the linear regime
for larger Γh values.
Here f mem is the surface force density arising from the
vesicle membrane. The tensor τ includes contributions
from both the hydrodynamic and electrostatic stresses:
τ ≡ −pI + µ(∇v + ∇vT ) + ǫEE −
1
2
ǫ(E · E)I.
(20)
verify that f κ is several orders of magnitude smaller than
f Γ, and is therefore not included in the current analysis.
The local membrane tension, Γ, is calculated by assum-
ing an effective tension which is uniform over the entire
membrane.6,29 An increase of the homogeneous tension,
Γh, from the initial tension, Γ0, leads to an increase in
the apparent membrane area:1,29 -- 31
∆ =
kBT
8πκ
ln
Γh
Γ0
+
Γh − Γ0
Ka
.
(22)
Here ∆ is the increase in the apparent membrane area
relative to the initial spherical state,
1
∆ =
0 (cid:1)−
2(cid:0)1 − ξ−2
2
3 h1 − ξ−2
0 +(cid:0)ξ2
0 − 1(cid:1)
1
2 arcsin(cid:0)ξ−1
0 (cid:1)i−1.
(23)
Ka is the elastic stretching modulus. κ is the bending
rigidity. Equation (22) indicates that Γ0, κ, and Ka are
the important parameters in determining membrane ten-
sion. κ and Ka are usually constants for a specific vesicle
type, and their values are often readily obtained from
previous work.1,32,33 On the other hand, Γ0 is specific to
an individual vesicle, and its value can not be directly
determined from experimental measurements. The re-
lation between ∆ and Γh for different choices of Γ0 is
shown in Fig. 2. When ∆ is small, the membrane area
increases through the flattening of the undulations, and
Γh shows an exponential correlation with ∆. When ∆ is
sufficiently large, a linear behavior is observed instead,
and the membrane area increase is mainly due to elastic
stretching. Moreover, a larger Γ0 always leads to a larger
Γh for the same value of ∆.
C. The membrane-mechanical model
The surface force density at the vesicle membrane es-
sentially consists of two parts6,28
f mem = f κ + f Γ.
(21)
Here f κ is the surface force density induced by bending
resistance. f Γ = 2ΓH n−∇sΓ is the surface force density
induced by the membrane tension. H is the mean curva-
ture, and Γ is the local membrane tension. We can easily
D. General solution
A solution for vesicle electrodeformation can be ob-
tained by solving the governing equations of both the
electrical and hydrodynamic problems, with the help of
the matching conditions. The solution strategy is iden-
tical to that presented in Zhang et al.,13 with only dif-
ferences in the detailed matching conditions for both the
electric field and the interfacial forces. For brevity, only
the final governing equation for ξ0 is presented here:
dξ0
dτ
= −
QN =
CaE
1
Γh
Γ0
µrf22(ξ0) + f23(ξ0)
µrf14(ξ0) + f15(ξ0) −
F (cid:20)QN f21(ξ0) + QT
λ2 h(λ − αQ
λ2 h(λ − αQ
f24(ξ0)(cid:21) ,
1(ξ0))2 + (λ − αQ1(ξ0)/ξ0)2 − 2β2/ǫri ,
1(ξ0))(λ − αQ1(ξ0)/ξ0) − β2/ǫri .
CaE
′
′
QT =
The functions f14(ξ0), f15(ξ0), f21(ξ0) − f24(ξ0), and F
are the same as those used in Zhang et al.,13 and the
detailed expressions are found in the Appendix.
(24a)
(24b)
(24c)
ǫr ≡
Transient solution for vesicle electrodeformation and relaxation
5
ǫe/ǫi is the permittivity ratio. The factors QN and QT
again arise from the effects of the tangential and normal
stresses, respectively. CaE ≡ r0ǫeE2
0 /Γ0 is the modified
electric capillary number.
In the absence of electropo-
ration, the coefficients α and β are given in Eqs. (11)
and (12). Once the electroporation occurs, Eq. (14) is
used instead. Similar to the droplet model, an examina-
tion of the three terms in the numerator of Eq. (24a)
reveals the contribution from the normal stress, tangen-
tial stress, and membrane tension, respectively. The bal-
ance between these three terms determines the equilib-
rium vesicle shape. The above equations are solved until
the end of the pulse, t = tp.
In the context of vesicle electrodeformation, the relax-
ation process is equally important, and is more reveal-
ing of the underlying physical processes. The governing
equations are presented below. In the absence of electro-
poration, Eq. (3) is solved without an applied electric
field. The resulting equation for ξ0 remains the same as
Eq.
(24a). The coefficients of QN , QT , α, and β are
given as
QN =
ǫeV 2
c
λ2r0Γ0 hα2(cid:16)Q
′2
1 (ξ0) + Q2
1(ξ0)/ξ2
0(cid:17) − 2β2/ǫri ,
(25)
QT =
ǫeV 2
c
λ2r0Γ0 hα2Q1(ξ0)Q
′
1(ξ0)/ξ0 − β2/ǫri ,
α =
β
Q′
1(ξ0)σr
,
dξ0
dτ
+
τ2
τ1λ# β = 0,
(cid:20) Q1(ξ0)
1(ξ0)σr − ξ0(cid:21) dβ
Q′
dτ −" Q1(ξ0)Q
′′
1 (ξ0) − Q
1 (ξ0)(1 − σr)
′2
Q′2
1 (ξ0)σr
α(τp) =
Vc(Q′
Vm(τp)
1(ξ0)ξ0σr − Q1)
,
β(τp) =
Vm(τp)Q
′
1(ξ0)σr
Vc(Q′
1(ξ0)ξ0σr − Q1)
.
(26)
(27)
(28)
(29)
(32)
(33)
(34)
In Eq. (29), the initial conditions for α and β are ob-
tained by solving Eqs. (3) and (6), and requiring that
Vm assumes the value at the end of the pulse. τp is the
dimensionless time, tp/τ2. Note that in this case, al-
though the pulse is switched off, the electric field is in
general not zero, due to the capacitive discharging of the
membrane. In this case, the TMP will decreases from its
peak value to zero on the membrane-charging timescale,
Tch.
When electroporation is present, the discharging pro-
cess is slightly more complex. The full membrane-
charging model (5) is used.
In order to determine the
membrane conductance, Gm, we simply assume that it
remains unchanged from the moment the pulse ceases,
namely,
Gm = −
σeβE0
λVc
.
(30)
The resulting equation for ξ0 again does not formally
deviate from Eq. (24a). The coefficients of QN , QT , α,
and β are
QN =
ǫeV 2
c
λ2r0Γ0 hα2(cid:16)Q
′2
1 (ξ0) + Q2
1(ξ0)/ξ2
0(cid:17) − 2β2/ǫri ,
(31)
QT =
ǫeV 2
c
λ2r0Γ0 hα2Q1(ξ0)Q
′
1(ξ0)/ξ0 − β2/ǫri ,
β
α =
,
Q′
1(ξ0)σr
(cid:20) Q1(ξ0)
1(ξ0)σr − ξ0(cid:21) dβ
Q′
dτ −" Q1(ξ0)Q
′′
1 (ξ0) − Q
1 (ξ0)(1 − σr)
′2
Q′2
1 (ξ0)σr
dξ0
dτ
+
τ2
τ1λ −
τ2Gm
Cm (cid:18) Q1(ξ0)
1(ξ0)σr − ξ0(cid:19)# β = 0.
Q′
E. A similarity solution for vesicle relaxation
The governing equation for the relaxation process can
be further simplified following two considerations. First,
we may ignore the membrane-discharging process. The
membrane-charging/discharging time, Tch, is on the or-
der of 1 ms, which is in general much shorter than the
Transient solution for vesicle electrodeformation and relaxation
6
relaxation time observed in the experiments, namely, a
few tens of ms or longer. The relatively small effect of
discharging on relaxation is clearly seen in Fig. 3 pre-
sented in the following section. Without including the
discharging process, the coefficients QT and QN in Eq.
(24a) are simply set to zero. Second, in the membrane-
mechanical model (22), the first and second term on
the RHS represent the effects of undulation unfolding
and elastic stretching, respectively. For moderate values
of Γ0, and for small-to-moderate deformations, the sec-
ond term can be ignored, and the membrane-mechanical
model becomes
∆ =
kBT
8πκ
ln
Γh
Γ0
.
(35)
Substituting QT = QN = 0 and Eq. (35) into (24a), we
obtain
dξ0
dτ
=
1
F
exp(
8πκ∆
kBT
)f24(ξ0).
(36)
This equation is conveniently rewritten in terms of the
aspect ratio as
d a
b
dτ
1
F
= −
exp(
8πκ∆
kBT
)(ξ2
0 − 1)−
3
2 f24(ξ0).
(37)
Note that in this equation, κ, the bending rigidity, is
regarded constant for a specific vesicle type, and µr (em-
bedded in F , see Appendix) is close to 1 as both the fluids
are usually aqueous. In addition, ∆, the relative increase
of apparent membrane area, depends exclusively on ξ0,
hence a/b according to Eqs. (23) and (2). Under these
assumptions, we observe that Eq. (37) is completely au-
tonomous, and the relaxation process is governed by the
dimensionless time, τ = t/τ2, where τ2 = r0µe/Γ0. This
result suggests that the relaxation of vesicles with dif-
ferent initial radius, r0, and initial tension, Γ0, obeys
a similarity behavior with the proper scaling suggested
above. This behavior is demonstrated by both simulation
and analysis of previous experimental data below.
III. RESULTS
For all results below, we assume the lipid membrane
to be made of egg-PC following Riske and Dimova8
(henceforth abbreviated as ′RD05′) and Sadik et al.10
(henceforth denoted as ′S11′). The bending rigidity is
taken to be κ = 2.47 × 10−20 J;1 the elastic modu-
lus, Ka = 0.14 N/m;32,33 the membrane capacitance,
Cm = 0.01 F/m2;34 the intravesicular and extravesicular
viscosities, µi = µe = 10−3 Pa · s; the intravesicular and
extravesicular permittivities, ǫi = ǫe = 7 × 10−10 F/m.
The critical transmembrane potential is assumed to be
Vc = 1 V.35
TABLE I. List of parameters for Fig. 5. For each case, E0 and
tp are specified according to RD05. Γ0 is a fitting parameter
to obtain best comparison between simulation and data. For
cases b, d, e, and f, extended pulse lengths (denoted by star)
are also used.
case # E0 (kV/cm)
tp (µs)
a
b
c
d
e
f
g
h
1
1
1
1
1
1
2
2
2
2
3
3
150
200
300*
250
300
400*
50
80*
100
170*
50
100
Γ0 (N/m)
2.79 × 10−4
3.23 × 10−6
3.23 × 10−6
1.67 × 10−4
1.80 × 10−6
1.80 × 10−6
1.80 × 10−4
1.80 × 10−4
3.16 × 10−6
3.16 × 10−6
6.67 × 10−6
3.42 × 10−7
A. The effects of Γ0 and tp
We begin by examining the effects of Γ0 on vesi-
cle electrodeformation and relaxation. Figure 3 shows
the typical system behavior for values of Γ0 ranging
from 10−7 − 10−3 N/m. The intravesicular and ex-
travesicular conductivities are σi = 6 × 10−4 S/m and
σe = 4.5 × 10−4 S/m, respectively following RD05. The
field strength is E0 = 1 kV/cm, the pulse length is
tp = 250 µs, and the initial radius is r0 = 15 µm. Fig-
ure 3(a) shows the evolution of Vm at the cathode-facing
pole, which demonstrates only a weak dependence on Γ0.
The threshold for electroporation (1 V) is reached just
before the end of the pulse, and its effects are present yet
negligible. The discharging occurs on the relatively short
timescale of 1 ms as we discussed above. Figure 3(b)
shows the evolution of the aspect ratio, a/b. The dis-
charging process manifests itself as a sudden and slight
decrease in the aspect ratio immediately after the pulse
ceases; its effects can in general be ignored without signif-
icantly altering the relaxation behavior. A smaller value
of Γ0 leads to a larger aspect ratio, and a longer relax-
ation process. The maximum aspect ratio, [a/b]max, is
plotted as a function of Γ0 in Fig. 3(c). As the initial
membrane tension decreases toward zero, the maximum
achievable aspect ratio saturates.
The similarity behavior in the relaxation process is
demonstrated in Fig. 3(d). The descending branches
of the curves (t > tp) shown in Fig. 3(b) are rescaled in
terms of τ = t/τ2, and shifted horizontally. In compari-
son, the thick solid curve is obtained by directly solving
Eq. (37). The convergence of all curves validates that
τ2 = r0µe/Γ0 is the single timescale governing vesicle
relaxation.
The effects of tp are examined in Fig. 4. The pa-
rameters are the same as in Fig. 3, and we fix Γ0 at
1 × 10−6 N/m. Figure 4(a) shows that a longer pulse
Transient solution for vesicle electrodeformation and relaxation
7
Discharging
Γ0 = 1.0 × 10−3 N/m
Γ0 = 1.5 × 10−4 N/m
Γ0 = 1.0 × 10−5 N/m
Γ0 = 1.0 × 10−6 N/m
Γ0 = 1.0 × 10−7 N/m
(b)
1.2
1.15
a b
1.1
1.05
(a)
)
V
(
m
V
1
0.8
0.6
0.4
0.2
0
0
(c)
1.2
x
a
m
]
a b
[
1.15
1.1
1.05
0.2
0.4
0.6
0.8
1
t (ms)
1
10−6 10−5 10−4 10−3 10−2 10−1 100 101
t (s)
Similarity Solution
(d)
1.2
1.15
a b
1.1
1.05
1
10−7
10−6
10−5
10−4
10−3
Γ0 (N/m)
1
0
2
4
τ
6
8
10
FIG. 3. Vesicle deformation-relaxation as a function of Γ0. The governing parameters are σi = 6×10−4 S/m, σe = 4.5×10−4 S/m,
E0 = 1 kV/cm, tp = 250 µs, and r0 = 15 µm. (a) The transmembrane potential at the cathode-facing pole. (b) The time-course
of the aspect ratio. (c) The maximum aspect ratio as a function of Γ0. (d) The similarity behavior in relaxation. The descending
branches from (b) are rescaled with τ = t/τ2. The thick solid curve is directly obtained by integrating Eq. (37).
consistently leads to greater deformation, and the aspect
ratio increases along the same envelope. The relaxation
times are approximately the same for all cases, because
τ2 remains unchanged. The discharging process is in gen-
eral more conspicuous with longer pulses. In Fig. 4(b),
the relaxation curves are again shifted horizontally and
rescaled with τ2 to show good agreement with the simi-
larity solution (thick solid line). Note that here because
all cases share the same values of τ2, the collapse of the
curves is primarily caused by simple shifting.
In other
words, the aspect ratio also decreases along a common
envelope.
The above results are exemplary and demonstrate the
typical system behavior. In general, the relaxation pro-
cess (in particular the relaxation time) is more apprecia-
bly affected by the change in Γ0 than the deformation
process. A wide range of pulsing parameters are studied
below, in direct comparison with experimental data from
RD05 and S11.
B. Comparison with experimental data
An extensive comparison of our theoretical prediction
with the data from RD05 is presented in Fig. 5. For
all eight cases, the initial radius is r0 = 15 µm. The
electrical conductivities are σi = 6 × 10−4 S/m and
σe = 4.5 × 10−4 S/m, respectively, leading to a conduc-
tivity ratio of σr = 0.75. Other parameters are listed in
table I. All parameters are taken directly from RD05, ex-
cept for the extended pulse lengths for some cases noted
below. For each case, the initial tension, Γ0, is deter-
Transient solution for vesicle electrodeformation and relaxation
8
(a)
1.25
a b
1.2
1.15
1.1
1.05
tp = 50 µs
tp = 100 µs
tp = 150 µs
tp = 200 µs
tp = 250 µs
tp = 300 µs
1
10−6 10−5 10−4 10−3 10−2 10−1 100 101
t (s)
Similarity Solution
(b)
1.25
a b
1.2
1.15
1.1
1.05
1
0
2
4
τ
6
8
10
FIG. 4. Vesicle deformation-relaxation as a function of tp.
The parameters are the same as in Fig. 3. The initial tension
is set to be constant, Γ0 = 1 × 10−6 N/m. (a) The time-course
of the aspect ratio. (b) The similarity behavior is observed
by shifting the relaxation curves with respect to time. The
relaxation timescale, τ2 = r0µe/Γ0, is the same for all cases.
The thick solid curve is directly obtained by integrating Eq.
(37).
mined to best fit the experimental data; their values are
listed in table I in the last column. The experimental
data are presented as symbols; the theoretical predic-
tions, solid lines. In Figs. 5(a) to 5(d), the electric field
strength is E0 = 1 kV/cm. For these cases, Vm is pre-
dicted to reach Vc at t = 242 µs.
In Figs. 5(a) and
5(c), good agreements are observed between the theoret-
ical prediction and the data. In Figs. 5(b) and 5(d), the
model results underpredict the maximum aspect ratios.
This discrepancy is peculiar: our simulation follows the
data accurately during the presence of the pulse, which
duration is provided by RD05. After the pulse ceases,
the simulation predicts immediate relaxation, whereas
the vesicles continued to deform in the experiments, due
to some unknown cause. In an attempt to mend this dif-
ference, we artificially increase the pulse lengths in the
simulation in b and d from 200 and 300 to 300 and 400 µs,
respectively. The values for Γ0 remain unchanged. The
results are shown as dashed curves. The model predicts
well the data for both the deformation and relaxation
processes. Note that although the relaxation curves rep-
resented by the solid and dashed lines look somewhat
different due to the semi-log scale on the time axis, they
actually follow the same descending envelopes which we
have demonstrated in Fig. 4(b) above.
In Figs. 5(e) and 5(f), the field strength is increased to
be E0 = 2 kV/cm, and the pulse lengths used in RD05
were 50 and 100 µs, respectively. For these cases, our
model predicts the occurrence of electroporation around
t = 103 µs. A similar situation is observed as in Figs.
5(b) and 5(d). The solid curves underpredict the maxi-
mum aspect ratio. Artificially extending the pulses in e
and f to 80 and 170 µs, respectively, leads to much better
agreement between the two.
In Figs. 5(g) and 5(h), the field strength is further
increased to 3 kV/cm, and electroporation is predicted
to occur at t = 66 µs. The entire deformation-relaxation
process is well-captured in g where tp = 50 µs. In Fig.
5(h), where tp = 100 µs, although the model accurately
predicts the deformation, the simulated relaxation curve
completely deviates from the experimental data. For this
case, and for pulses even longer than 100 µs, RD05 [Fig.
1(c) therein] exhibits a regime where complex, multi-
stage relaxation process was observed.
In this regime,
the membrane structure is likely severely altered due to
electroporation, which process can not be captured by
our present model. Further comparison with these data
is not pursued.
The similarity behavior in the relaxation process is
demonstrated in Fig. 6. The experimental data from
Figs. 5(a) to 5(g) are shifted horizontally and rescaled
with τ2. For each case, τ2 is obtained using Γ0 listed in
table I. The thick solid curve is again the similarity so-
lution from Eq. (37), and the results are shown on both
semi-log and linear scales in τ . The coefficient of de-
termination is R2 = 0.96. The experimental data from a
wide range of parameters demonstrate a universal behav-
ior governed by a single timescale, τ2 = r0µe/Γ0. This
result is a main contribution of the present work.
We remark that a similar behavior should be observed
for droplets, where the initial membrane tension, Γ0, is
replaced by γ, the coefficient of surface tension in τ2 (cf.
the definition of τ2 in Zhang et al.).13 However, there is
a subtle difference between droplet and vesicle relaxation
while the coefficient of surface tension is usually a con-
stant, the membrane tension, Γh, is not. Nonetheless,
as long as Γh depends linearly on Γ0, which is a good
approximation for small-to-moderate deformations. The
universal behavior in Fig. 6 is expected.
Finally, the model prediction is compared with data
Transient solution for vesicle electrodeformation and relaxation
9
(b)
1.25
1.2
1.15
1.1
1.05
10−5
10−4
1
10−6
10−5
10−4
10−3
10−2
(d)
1.4
10−1
(a)
1.08
1.06
a b
1.04
1.02
1
10−6
(c)
1.12
a b
1.1
1.08
1.06
1.04
1.02
1
0.98
10−6
(e)
1.15
a b
1.1
1.05
1
10−5
(g)
1.25
a b
1.2
1.15
1.1
1.05
10−5
10−4
10−3
10−4
1.3
1.2
1.1
1
0.9
10−6
1.5
(f )
10−5
10−4
10−3
10−2
10−1
1.4
1.3
1.2
1.1
1
0.9
10−6
(h)
1.6
10−5
10−4
10−3
10−2
10−1
1.5
1.4
1.3
1.2
1.1
1
10−6
10−5
10−4
10−3
10−2
t (s)
1
10−6
10−5
10−4
t (s)
10−3
10−2
FIG. 5. Comparison with the deformation-relaxation data from RD05. For all cases, r0 = 15 µm, σi = 6 × 10−4 S/m, and
σe = 4.5 × 10−4 S/m. Parameters specific to each case are listed in table I. The data is represented by symbols, and the
simulation is represented by solid curves. For cases b, d, e, and f, the dashed lines represent the simulated results with extended
pulses (denoted by stars in table I).
(a)
1.4
a b
1.3
1.2
1.1
1
0.9
10−2
(b)
1.4
a b
1.3
1.2
1.1
1
10
σr = 5.00 × 10−1
σr = 1.67 × 10−2
σr = 3.13 × 10−2
σr = 2.17 × 10−2
σr = 1.89 × 10−2
100
200
300
400
500
t (µs)
Current Model
S11
20
1
σr
40
60
2
1.8
a b
1.6
1.4
1.2
1
0
(b)
2.5
a b
2
1.5
1
0
Transient solution for vesicle electrodeformation and relaxation
Similarity Solution
(a)
2.2
10−1
τ
100
101
Similarity Solution
0.9
0
2
4
τ
6
8
10
FIG. 6. The similarity behavior of vesicle relaxation. The
experimental data from cases a-g in Fig. 5 are shifted in
time, then rescaled by τ2 = r0µe/Γ0. They are represented
by symbols. The solid curves are calculated with Eq. (37).
The same data are shown on both a semi-log (a) and a linear
(b) scale. The coefficient of determination is R2 = 0.96.
from S11.
In this work, the deformation is examined
at a fixed pulse length of tp = 500 µs, and for five
intra-to-extra vesicular conductivity ratios. Only the
case of E0 = 0.9 kV/cm is examined, where no or weak
electroporation is expected. We do not compare the
cases of E0 = 2 and 3 kV/cm in S11, where the vesi-
cles were in the strongly-electroporated regime, and our
model no longer applies. The governing parameters are
r0 = 11.3 µm and σe = 3 × 10−4 S/m. The initial mem-
brane tension is chosen to be the same for all vesicles,
namely, Γ0 = 1×10−8 N/m. Figure 7(a) shows the defor-
mation process as a function of time for five conductivity
ratios. As σr decreases the rate of deformation increases.
Except for the case of σr = 0.5, the aspect ratio reaches
FIG. 7. Comparison with data from S11. (a) Simulated time-
course of the aspect ratio for various conductivity ratios. For
all cases r0 = 11.3 µm and Γ0 = 1 × 10−8 N/m. (b) The
aspect ratio at t = 500 µs as a function of 1/σr.
a plateau before the pulse ends. The time at which the
aspect ratio increases saturates with an increasing σr.
For σr = 0.5, an equilibrium could be reached if the
pulse length is extended and sufficiently long (not shown
here). In Fig. 7(b), the aspect ratio at t = tp is shown
as a function of 1/σr. We choose this representation to
facilitate comparison with the data from S11 (symbols),
where the definition of the conductivity ratio is σi/σe.
A reasonable agreement is found between the two. The
behavior of the simulation and the data is explained by
the dependence of the electrical stress on σr in S11 [see
Eq. (21) and Sec. 4 therein]. We do not repeat it here
for brevity. The current model represents a significant
improvement from that in S11, where the hydrodynamic
problem is treated empirically.
Some remarks are appropriate before concluding the
Transient solution for vesicle electrodeformation and relaxation
11
section. First, for most cases studied here, the TMP is
near the threshold, and the vesicles are expected to ex-
perience no or weak electroporation. For this regime,
our model is shown to provide a good predictive capabil-
ity, which demonstrates that the membrane-mechanical
model (22), although derived assuming no electropo-
ration, can be extended to the weakly-electroporated
regime, presumably due to the absence of major struc-
tural alterations. Our model is not applicable to the
strongly-electroporated regime. Second, the universal
scaling law in relaxation observed in Figs. 3, 4, and 6 is
expected to hold regardless of the means of deformation,
e.g., via AC/DC electric fields, or via mechanical stretch-
ing. Equation (37) is applicable to a wide range of relax-
ation phenomena beyond electrodeformation. Third, the
current work suggests that an extensive parametric study
on vesicle electrodeformation-relaxation experimentally,
in particular in the sub-critical regime where electropora-
tion is avoided, can provide the benefit to further validate
our model understanding. A systematic approach can be
possibly developed based on this work to map membrane
properties.
IV. CONCLUSIONS
In this work, we developed a transient analysis for vesi-
cle electrodeformation. The theory is derived by extend-
ing our previous work on a droplet model in Zhang et
al.,13 with the additional consideration of a lipid mem-
brane separating two fluids of arbitrary properties. For
the latter, both a membrane-charging and a membrane-
mechanical model are supplied. Similar to the droplet
model, the main result is also an ODE governing the
evolution of the vesicle aspect ratio. The effects of initial
membrane tension and pulse length are examined. The
initial membrane tension affects the relaxation process
much more significantly than the deformation process,
in particular when its value is small. The model predic-
tion is extensively compared with experimental data from
Riske and Dimova8 and Sadik et al.,10 and is shown to
accurately capture the system behavior in the regime of
no or weak electroporation. More importantly, the com-
parison reveals that vesicle relaxation obeys a universal
behavior, and is governed by a single timescale that is a
function of the vesicle initial radius, the fluid viscosity,
and the initial membrane tension. This behavior is re-
gardless of the means of deformation, either via AC/DC
electric field, or via mechanical stretching. This universal
scaling law is a main contribution of the current work,
and can be used to calculate membrane properties from
experimental data.
ACKNOWLEDGMENTS
JZ and HL acknowledge fund support from an NSF
award CBET-0747886 with Dr William Schultz and Dr
Henning Winter as contract monitors.
Appendix A
The functions f14(ξ0), f15(ξ0), f21(ξ0) − f24(ξ0), and
F in Eq. (24a) are given in the following expressions:
dη,
(A1)
f11(ξ0) =Z G3(η)η
(ξ2
0 − η2)
0 − 1(cid:26)Z G3(η)η
(ξ2
1
ξ2
0 − η2)(cid:18) (1 − 3η2)
0 − η2) − 3(cid:19) dη(cid:27) ,
(ξ2
(A2)
′′
G
3 (ξ0)G
′
5(ξ0) − G
2N
′
′′
3(ξ0)G
5 (ξ0)
· f11(ξ0), (A3)
f12(ξ0) =
f13(ξ0) =
f14(ξ0) = −ξ0H
′
3(ξ0)Z G3(η)η
0 − η2)2 dη +
(ξ2
1
2
H
′′
3 (ξ0)f11(ξ0),
(A4)
H
f15(ξ0) = −
′
3(ξ0)hG3(ξ0)G
′′
5 (ξ0) − G
2N
′′
3 (ξ0)G5(ξ0)i
f11(ξ0) + ξ0H
′
3(ξ0)Z G3(η)η
0 − η2)2 dη.
(ξ2
(A5)
Here N ≡ G3(ξ0)G
The detailed expressions for G and H are found in Dassios et al..36
5(ξ0)− G
3(ξ0)G5(ξ0). G and H are Gegenbauer functions of the first and second kind, respectively.
′
′
f21(ξ0) =
1
2
ξ2
dη,
0Z (η2 − 1)(3η2 − 1)
0 − 3ξ2
(ξ2
0 − η2)2
(ξ2
0 − η2)
0η2 + 2η4)
′
f22(ξ0) = ξ0f11(ξ0)(cid:20)−H
f23(ξ0) = ξ0f11(ξ0)"−
3(ξ0)Z (1 − 3η2)(ξ2
49(1 − 3ξ2
0)G3(ξ0)H
30N
′
dη + 3ξ0H3(ξ0)Z 1 − 3η2
(ξ2
0 − η2)
3(ξ0)
+ H
′
3(ξ0)Z (1 − 3η2)(ξ2
0 − 3ξ2
(ξ2
0 − η2)2
0η2 + 2η4)
(A6)
(A7)
(A8)
dη(cid:21) ,
dη# ,
Transient solution for vesicle electrodeformation and relaxation
f24(ξ0) = ξ3
0(1 − ξ−2
0 )
5
6 Z
3η2 − 1
(ξ2
0 − η2)
dη + ξ0(1 − ξ−2
0 )−
3
2
1
6 Z
3η2 − 1
pξ2
0 − η2
dη,
F = −
2
3
(f25(ξ0) + f26(ξ0)/µr) ,
where
f25(ξ0) = −
f22(ξ0)
ξ0f11(ξ0)
(µr − 1)f12(ξ0) + f13(ξ0)
µrf14(ξ0) + f15(ξ0)
f26(ξ0) = −
f23(ξ0)
ξ0f11(ξ0)
(µr − 1)f12(ξ0) + f13(ξ0)
µrf14(ξ0) + f15(ξ0)
µr ≡ µe/µi is the viscosity ratio.
− 3ξ0Z 3η2 − 1
(ξ2
0 − η2)
49(1 − 3ξ2
30N
0)G
−
′
dη −
ξ2
3(ξ0)
+
ξ2
ξ0
0 − 1Z (2ξ2
0 − 1Z (2ξ2
ξ0
0 − η2 − 1)(1 − 3η2)2
(ξ2
0 − η2)2
0 − η2 − 1)(1 − 3η2)2
(ξ2
0 − η2)2
12
(A9)
(A10)
dη, (A11)
dη. (A12)
1M. Kummrow and W. Helfrich, "Deformation of giant lipid vesi-
cles by electric fields," Phys. Rev. A 44, 8356 -- 8360 (1991).
2G. Niggemann, M. Kummrow, and W. Helfrich, "The bending
rigidity of phosphatidylcholine bilayers: Dependences on exper-
imental method, sample cell sealing and temperature," J. Phys.
II France 5, 413 -- 425 (1995).
3P. K. Wong, W. Tan, and C. M. Ho, "Cell relaxation after elec-
trodeformation: effect of latrunculin a on cytoskeletal actin,"
Biomech. J. 38, 529 -- 535 (2005).
4R. Dimova, K. A. Riske, A. S., N. Bezlyepkina, R. L. Knorr,
and R. Lipowsky, "Giant vesicles in electric fields," Soft Matter
3, 817 -- 827 (2007).
5S. Aranda, K. A. Riske, R. Lipowsky, and R. Dimova, "Mor-
phological transitions of vesicles induced by alternating electric
fields," Biophys. J. 95, L19 -- L21 (2008).
6P. M. Vlahovska, R. S. Gracia, S. Aranda-Espinoza, and R. Di-
mova, "Electrohydrodynamic model of vesicle deformation in al-
ternating electric fields," Biophys. J. 96, 4789 -- 4803 (2009).
7S. Kakorin, T. Liese, and E. Neumann, "Membrane curvature
and high-field electroporation of lipid bilayer vesicles," J. Phys.
Chem. B 107, 10243 -- 10251 (2003).
8K. A. Riske and R. Dimova, "Electro-deformation and poration
of giant vesicles viewed with high temporal resolution," Biophys.
J. 88, 1143 -- 1155 (2005).
9K. A. Riske and R. Dimova, "Electric pulses induce cylindrical
deformations on giant vesicles in salt solutions," Biophys. J. 91,
1778 -- 1786 (2006).
10M. M. Sadik, J. Li, J. W. Shan, D. I. Shreiber, and H. Lin, "Vesi-
cle deformation and poration under strong dc electric fields,"
Phys. Rev. E 83, 066316 (2011).
11H. Hyuga, K. J. Kinosita, and N. Wakabayashi, "Deformation
of vesicles under the influence of strong electric fields," Jpn. J.
Appl. Phys. 30, 1141 -- 1148 (1991).
12H. Hyuga, K. J. Kinosita, and N. Wakabayashi, "Deformation
of vesicles under the influence of strong electric fields II," Jpn. J.
Appl. Phys. 30, 1333 -- 1335 (1991).
13J. Zhang, J. D. Zahn,
and H. Lin, "A transient solution
for droplet deformation under electric fields," arXiv:1210.7878
[physics.flu-dyn] (2012).
14H. P. Schwan, "Dielectrophoresis and rotation of cells," in Elec-
troporation and electrofusion in cell biology, edited by E. Neu-
mann, A. E. Sowers, and C. A. Jordan (Plenum Press, 1989).
15C. Grosse and H. P. Schwan, "Celluar membrane potentials in-
duced by alternating fields," Biophys. J. 63, 1632 -- 1642 (1992).
16K. A. Debruin and W. Krassowska, "Modeling electroporation
in a single cell. I. Effects of field strength and rest potential,"
Biophys. J. 77, 1213 -- 1224 (1999).
17W. Krassowska and P. D. Filev, "Modeling electroporation in a
single cell," Biophys. J. 92, 404 -- 417 (2007).
18J. Li and H. Lin, "Numerical simulation of molecular uptake via
electroporation," Bioelectrochemistry 82, 10 -- 21 (2011).
19D. C. Chang and T. S. Reese, "Changes in membrane structure
induced by electroporation as revealed by rapid-freezing electron
microscopy," Biophys. J. 58, 1 -- 12 (1990).
20H. Leontiadou, A. E. Mark, and S. J. Marrink, "Molecular dy-
namics simulations of hydrophilic pores in lipid bilayers," Bio-
phys. J. 86, 2156 -- 2164 (2004).
21A. A. Gurtovenko and I. Vattulainen, "Pore formation coupled
to ion transport through lipid membranes as induced by trans-
membrane ionic charge imbalance: atomistic molecular dynamics
study," J. Am. Chem. Soc. 127, 17570 -- 17571 (2005).
22M. Tarek, "Membrane electroporation: a molecular dynamics
simulation," Biophys. J. 88, 4045 -- 4053 (2005).
23J. Wohlert, W. K. den Otter, O. Edholm, and W. J. Briels,
"Free energy of a trans-membrane pore calculated from atomistic
molecular dynamics simulations," J. Chem. Phys. 124, 154905
(2006).
24U. Pliquett, R. P. Joshi, V. Sridhara, and K. H. Schoenbach,
"High electrical field effects on cell membranes," Bioelectrochem-
istry 70, 275 -- 282 (2007).
25M. L. Fernandez, G. Marshall, F. Sagues,
and R. Reigada,
"Structural and kinetic molecular dynamics study of electropo-
ration in cholesterol-containing bilayers," J. Phys. Chem. B 114,
6855 -- 6865 (2010).
26N. Dubash and A. J. Mestel, "Behaviour of a conducting drop in
a highly viscous fluid subject to an electric field," J. Fluid Mech.
581, 469 -- 493 (2007).
27N. Bentenitis and S. Krause, "Droplet deformation in dc electric
fields: the extended leaky dielectric model," Langmuir 21, 6194 --
6209 (2005).
28U. Seifert, "Configurations of fluid membranes and vesicles,"
Adv. Phys. 46, 13 -- 137 (1997).
29W. Helfrich and R. M. Servuss, "Undulations, steric interaction
and cohension of fluid membranes," Il Nuvo Cimento 3D, 137 --
151 (1984).
30E. Evans and W. Rawicz, "Entropy-driven tension and bending
elasticity in condensed-fluid membranes," Phys. Rev. Lett. 64,
2094 -- 2097 (1990).
31E. Evans, "Entropy-driven tension in vesicle membranes and un-
binding of adherent vesicles," Langmuir 7, 1900 -- 1908 (1991).
32R. Kwok and E. Evans, "Thermoelasticity of large lecithin bilayer
vesicles," Biophys. J. 35, 637 -- 652 (1981).
33D. Needham, "Cohesion and permeability of lipid bilayer vesi-
cles," in Permeability and Stability of Lipid Bilayers, edited by
E. A. Disalvo and S. A. Simon (CRC Press, Boca Raton, FL,
Transient solution for vesicle electrodeformation and relaxation
13
1995) pp. 49 -- 76.
34D. Needham and R. M. Hochmuth, "Electro-mechanical perme-
abilization of lipid vesicles. Role of membrane tension and com-
pressibility," Biophys. J. 55, 1001 -- 1009 (1989).
35T. Portet and R. Dimova, "A new method for measuring edge
tensions and stability of lipid bilayers: effect of membrane com-
position," Biophys. J. 99, 3264 -- 3273 (2010).
36G. Dassios, M. Hadjinicolaou, and A. C. Payatakes, "Gener-
alized eigenfunctions and complete semiseparable solutions for
Stokes flow in spheroidal coordinates," Q. Appl. Math. 52, 157 --
191 (1994).
|
1806.08066 | 1 | 1806 | 2018-06-21T04:46:54 | Dimensionality-dependent crossover in motility of polyvalent burnt-bridges ratchets | [
"physics.bio-ph",
"cond-mat.stat-mech",
"q-bio.SC"
] | The burnt-bridges ratchet (BBR) mechanism is a model for biased molecular motion whereby the construct destroys track binding sites as it progresses, and therefore acts as a diffusing forager, seeking new substrate sites. Using Monte Carlo simulations that implement the Gillespie algorithm, we investigate the kinetic characteristics of simple polyvalent BBRs as they move on tracks of increasing width. We find that as the track width is increased the BBRs remain nearly ballistic for considerable track widths proportional to the span (leg length) of the polyvalent walker, before transitioning to near-conventional diffusion on two-dimensional tracks. We find there exists a tradeoff in BBR track association time and superdiffusivity in the BBR design parameter space of span, polyvalency and track width. Furthermore, we develop an analytical model to describe the ensembleaverage motion on the track and find it is in good agreement with our Gillespie simulation results. This work offers insights into design criteria for de novo BBRs and their associated tracks, where experimentalists seek to optimize directionality and track association time. | physics.bio-ph | physics | Dimensionality-dependent crossover in motility of polyvalent burnt-bridges ratchets
Chapin S. Korosec, Martin J. Zuckermann, and Nancy R. Forde∗
Department of Physics, Simon Fraser University,
8888 University Drive, Burnaby, British Columbia, V5A 1S6, Canada
(Dated: June 22, 2018)
The burnt-bridges ratchet (BBR) mechanism is a model for biased molecular motion whereby
the construct destroys track binding sites as it progresses, and therefore acts as a diffusing forager,
seeking new substrate sites. Using Monte Carlo simulations that implement the Gillespie algorithm,
we investigate the kinetic characteristics of simple polyvalent BBRs as they move on tracks of
increasing width. We find that as the track width is increased the BBRs remain nearly ballistic
for considerable track widths proportional to the span (leg length) of the polyvalent walker, before
transitioning to near-conventional diffusion on two-dimensional tracks. We find there exists a trade-
off in BBR track association time and superdiffusivity in the BBR design parameter space of span,
polyvalency and track width. Furthermore, we develop an analytical model to describe the ensemble-
average motion on the track and find it is in good agreement with our Gillespie simulation results.
This work offers insights into design criteria for de novo BBRs and their associated tracks, where
experimentalists seek to optimize directionality and track association time.
I.
INTRODUCTION
Diffusion, driven by random thermal motion, results in
slow transport over long distances. Nature has overcome
this problem through the evolution of impressive protein-
based machines that achieve processive and directional
motion despite their noisy thermal environment. Within
the cell's cytoplasm, the molecular motors kinesin [1],
dynein [2], and myosin [3] achieve directional motion on
their intracellular tracks by converting chemical energy in
the form of ATP into mechanical stepwise translocation
[4, 5]. There are, however, other means by which cellular
systems can achieve directed motion besides conventional
cytoplasmic motors. In this work, we examine a class of
machines that achieve directional motion by a 'burnt-
bridges ratchet'(BBR) mechanism.
A BBR has a probability p of destroying a substrate
track site as it passes [6]. Upon a successful cleavage
event, the asymmetry produced in the track prevents
backwards stepping. Motion forwards is driven purely
by thermal motion without the need for an energetically
driven conformational change in the walker.
In order
to achieve processive motion the timescale of track as-
sociation must be long enough such that the BBR can
cleave the substrate, explore neighbouring sites, and rely
on thermal fluctuations to move. In one dimension, with
p = 1, the motion of a BBR is expected to be ballis-
tic, while in two dimensions the motion is expected to
resemble a self-avoiding walk. We also note that BBR
nanomachines can be considered as diffusing foragers,
where a parameter of interest is the number of cleavage
events before the walker depletes its local environment
and 'starves' (detaches) [7].
Matrix-metalloproteases (MMPs) are enzymes that
move one-dimensionally via a BBR mechanism along col-
lagen fibrils in the extracellular matrix [8, 9]. Individual
∗ [email protected], [email protected], [email protected]
MMPs have been observed to move superdiffusively along
their collagen tracks at speeds up to 5.8µm/s [10].
In
contrast to the one-dimensional motion of MMPs, the
protein-based ParA/ParB system found in bacteria is
an example of a two-dimensional BBR system [11, 12].
This system is responsible for partitioning extrachromo-
somal low-copy plasmid DNA during cell division [13].
These BBRs have been observed to move directionally
at speeds of ∼ 0.1 µm/s on their two-dimensional tracks
[11]. Nature has therefore implemented the BBR mech-
anism in both one-dimensional and two-dimensional sys-
tems, where these BBRs have achieved speeds compara-
ble to kinesin in saturating ATP conditions [14].
Inspiration from biological systems such as these
has led to the development of synthetic nanomachines
that achieve directional motion through various stepping
mechanisms [15–24]. The motivation for the design and
implementation of synthetic nanomachines is two-fold:
to create a molecular system that mimics the behaviour
of biological counterparts, thereby enabling us to learn
about fundamental physical principles that give rise to
observed biological molecular motor phenomena; and to
create new technologies that perform tasks currently out
of our reach [17].
Many of the autonomous synthetic biologically-based
nanomotors thus far realized are DNA-based BBRs
[16, 17, 20, 22].
In the limit of low polyvalency, Cha
et al. [22] developed a DNA walker that moves in a self-
avoiding fashion along carbon nanotubes by catalyzing
cleavage of its RNA footholds. At the other extreme,
DNA-coated microspheres, so-called 'DNA monowheels',
hybridize to a substrate surface coated with complemen-
tary RNA and have a high polyvalency with thousands of
cleavable substrate contacts [25]. The DNA monowheel
has demonstrated impressive velocities for an artificial
system of up to 2 µm/min, as well as near-ballistic mo-
tion on its two-dimensional substrate track [25]. In con-
trast to the monowheel's high polyvalency, most DNA
walkers are bipedal [19, 26, 27].
arXiv:1806.08066v1 [physics.bio-ph] 21 Jun 2018
In this work, we refer to the total number of legs as
the polyvalency. Polyvalency of BBRs is thought to have
a profound impact on directionality and track attach-
ment times [25]. Analytical approaches to understand-
ing the effects of polyvalency on BBR dynamics are dif-
ficult as the memory requirement for visited sites leads
to non-Markovian behaviour [28, 29]. Because of this,
researchers have largely turned to simulations to model
the behaviour of synthetic BBR nanomotors [30–33].
Those who seek to experimentally develop synthetic
nanomachines are met with the challenge of designing
not only the the machine itself, but also the substrate
track with which it is to interact. Missing from the liter-
ature is an exploration of how the width of the substrate
track is expected to impact BBR kinetics. In this work
we implement the Monte Carlo Gillespie algorithm [34] to
investigate the dependence of the mean squared displace-
ment, track attachment time, kurtosis, and extent of sub-
strate cleavage on the dimension of the substrate track.
We generalize our results by altering the polyvalency and
span of the BBRs to explore how these attributes influ-
ence ensemble-average kinetics of BBRs moving on tracks
of increasing width. In this work we focus on ideal BBRs
where substrate binding is followed by a probability p = 1
of catalyzing the bound site. Our BBRs cannot unbind
from substrate without a cleavage event, and cannot re-
bind to a cleaved product site. Samii et al.
[31] report
that nanomachines that can unbind from substrate and
rebind to product display an increase in track attachment
time as a function of increasing polyvalency. Similarly,
Yehl et al.
[25] report that the prolonged track attach-
ment of their BBR DNA monowheel is because of the
dramatically increased polyvalency. In contrast to these
results, in our system we find that increasing the poly-
valency of BBRs results in a dramatic decrease in track
association time. Our results further indicate that reduc-
ing the dimensionality of the track to one dimension is
not necessary to promote linear ballistic motion. There
exists a tolerance window in track width that allows for
maximally superdiffusive walkers.
2
for the unbound legs. We note that our model is funda-
mentally different from that of Olah et al. [32] where they
allowed binding to all sites within a distance ' of each
bound leg. We allow for binding within a small region
around the global constraint, where the region (shown in
yellow in Fig. 1a) is determined by the currently bound
legs. In this way we account for the collectively imposed
constraint of all bound legs limiting the options for fresh
track coupling.
To study the motion of BBRs we developed a kinetic
model similar to that used by Samii et al.
[30, 31] and
Olah et al. [32] whereby we implement the Monte Carlo
Gillespie algorithm [34] to study polyvalent walker dy-
namics. Our kinetic model, as shown in Fig. 1b, is a
simple model that allows for substrate binding and sub-
strate cleavage followed by unbinding. We employ a sub-
strate binding rate, kon = 20 s−1, and cleavage rate,
kef f = 0.054 s−1. kef f incorporates both the cleavage
and detachment processes. These rates are similar to
those used by Samii et al.
[30, 31]. We made the deci-
sion to set the dissociation rates from uncleaved substrate
sites to zero, and the product binding rate to zero. This
allows us to to focus on a strict burnt-bridges ratchet
system distinct from that of previous work on similar
systems [30–32]. Each substrate-bound leg is guaranteed
to cleave and release to the unbound state where it may
bind again to fresh substrate. Therefore, our legs have a
probability of p = 1 to cleave each bound substrate site.
A kinetic move is chosen by a Monte Carlo Gillespie
algorithm that samples from all available transitions of
all legs. For example, if there are 2 unbound legs, 3 bound
legs, and 12 available substrate sites, there are 24 possible
binding transitions and 3 possible transitions to cleave
and release. A particular transition with rate ki is chosen
. After a choice of transition
X ),
where X is a random variable uniformly distributed on
(0,1]. The central hub position is updated by determining
the average position of the bound legs. We then track the
motion of this point-like hub for kinetic analysis.
with a probability Pi = kiP ki
is made, time is updated according to t = 1P ki
ln( 1
II. MODEL AND METHODS
B. Track design & BBR parameters
A. Kinetic model
We model polyvalent BBRs as n legs coupled to a
point-like hub referred to as the global constraint (Fig.
1). The n legs are non-interacting, but only a single leg
can occupy any given track site. The legs chemically in-
teract with the track via substrate binding and cleavage,
followed by release from the cleaved product. To incorpo-
rate leg length into the kinetic model, each of the n legs
is assigned a span, which is defined as the maximum dis-
tance between any two bound legs. As shown in Fig. 1a
a circle of radius R = span is drawn around each bound
leg. The substrate track sites that fall within the mutual
overlap of all legs' spans are considered binding options
To explore the effect track dimensionality has on
BBRs, we employed a large range of tracks that increase
in width by factors of 2. In total we cover tracks of widths
2n for n = 0, 1, 2, 3, ..., 12, where a track width of 20 is a
one-dimensional track. For all 520,000 independent runs
reported in this work, track widths of 212 = 4096 can
be considered infinitely two-dimensional as no ratchet
reached the boundaries within the maximum simulation
time of 25,000 seconds for each independent run. As we
varied the track width, the length of the track was con-
sistently kept to 5000 lattice sites (effectively infinite in
length). The track widths chosen were convenient as they
allowed us to probe the effects of confinement through
gradually increasing the track size away from one dimen-
3
C. Analytical methods
1. Mean squared displacement
The mean squared displacement (MSD) is a useful
measure used to assess the anomalous nature of a dif-
fusive walk [35]. The MSD is defined as the variance
in displacement, ~X, and scales with a power-law depen-
dence,
MSD( ~X) ≡ Var[ ~X] =D( ~X − ~µ)2E ∝ tα,
(1)
where ~µ is the mean position and α describes the power-
law scaling behaviour of the system. For a given ensem-
ble of trajectories one can assess the slope of the log-log
MSD-time plots to compute α. For all BBR permuta-
tions and track widths we computed the MSD via en-
semble averaging according to eq. 2, which is equivalent
to eq. 1,
MSD(~r) = ((cid:10)x2(cid:11) − hxi2) + ((cid:10)y2(cid:11) − hyi2) ∝ tαr ,
where ~r is the position of the BBR's global constraint.
The x and y components of ~r can be examined indepen-
dently. We therefore also define MSD(x) and MSD(y)
as
(2)
MSD(x) =(cid:10)x2(cid:11) − hxi2 ∝ tαx ,
MSD(y) =(cid:10)y2(cid:11) − hyi2 ∝ tαy .
2. Kurtosis
(3)
(4)
Kurtosis is defined as the standardized fourth moment
of a distribution about its mean and is given by
Kurt[ ~X] =
(5)
m4
σ4 = D( ~X − ~µ)4E
(cid:16)D( ~X − ~µ)2E(cid:17)2 ,
where m4 is the fourth moment, σ the standard devia-
tion, and ~µ the mean. The kurtosis is a useful descriptor
for a distribution's deviation from Gaussian. Gaussian
distributions have a kurtosis of 3. Therefore, it is conve-
nient to define γ2, the excess kurtosis, as γ2 = m4
σ4 − 3.
As our BBRs progress on their respective tracks, they
produce time-evolving displacement distributions.
In
particular, we are interested in characterizing the dif-
ferences between the evolving x- and y- components of
the displacement distributions to understand the effect
of constraints on the dynamics of BBRs. To this end,
γ2 provides a measure of the shape of the distributions
and allows for easy comparison across a large parameter
space of BBR designs.
FIG. 1. a) Each bound leg is assigned a span which defines
the radius of a circle, shown in green, around the bound loca-
tion, shown in red. The mutual overlap of bound-leg spans,
shown in yellow, marks the feasible binding locations for un-
bound legs (not shown) during the following kinetic move. b)
Each BBR leg (red) can interact with any availble substrate
track site (grey triangle) via a rate of attachment, kon. kef f
includes both substrate cleavage and release.
sion into an infinite two-dimensional plane. All tracks
were initialized as all-substrate tracks; we do not impose
any initial asymmetry. All of the BBRs were initialized
in the geometric centre of the track with one leg bound.
Motivated by previous work [11, 12, 16, 20, 24, 25, 30–
33], we focused on four BBR designs. In the notation of
(polyvalency, span) we examined the behaviour of BBRs
with parameters (12,8), (3,8), (12,3), and (3,3). The limit
of (12,3) BBRs was chosen because all available binding
locations within a span of 3 can be saturated.
In this
limit the BBRs are expected to produce a complete wake
of cleaved sites such that the constructs cannot cross pre-
viously visited territory. This contrasts with (3,8) BBRs,
which we expected to produce sparsely cleaved trajecto-
ries.
For each BBR design, on each track width, we ran
10,000 independent trajectories. For example, (3,3) has
10,000 independent runs on a track width of 1, and an-
other 10,000 on a track width of 2, etc. Trajectories
end when no BBR legs remain coupled to the track. We
do not allow rebinding of detached BBRs. If the BBRs
remain attached to the track for 25,000 seconds the sim-
ulation is also ended.
(a)
Substrate site
Product site
Substrate-bound leg
Global constraint (hub)
Track sites within the span
(not feasible to bind)
Accepted binding sites
Span = 4
Span = 4
(b)
kon!
kef f!
III. RESULTS
A. Sample trajectories and distributions
Single trajectories in two dimensions for (3,8), (3,3),
(12,3) and (12,8) BBRs are presented in Fig. 2 (left
column). Our lowest span BBR systems, (12,3) and (3,3)
BBRs, tend to become easily entrapped by their product
wakes, leading to detachment. With increased span and
decreased polyvalency, such as the (3,8) system, BBRs
can reach over their previous trajectories into areas of
fresh substrate.
Snapshots of the ensemble behaviour of each BBR de-
sign in two dimensions are included in Fig. 2 (right col-
umn), beside their respective sample trajectories. The
(3,3), (12,3) and (12,8) systems all develop a low occu-
pancy near their centre starting position, while the (3,8)
system maintains the highest BBR occupancy around the
origin. These ensemble snapshots are taken from Movies
S1-S4, which present the full dynamic evolution of the
ensemble behaviour.
B. Mean squared displacement
Fig. 3a shows log-log plots of MSD(~r) vs. time for
(3,8) BBRs on all examined track widths. We report
values of α in the long-time limit when α ≈ 0. A one-
dimensional track results in ratchets moving ballistically
with α ≈ 2. As track width increases α begins to de-
crease. One would naıvely expect that as the width of
the track increases, the constructs have increased proba-
bility to change direction, thus lowering α. However, this
transition does not occur monotonically. Fig. 3b shows
that we observe a minimum in αr as a function of track
width for all BBR designs.
The non-monotonic behaviour of αr as a function of
track width prompted closer inspection of the MSD.
The log-log plot of MSD(x) (Fig. 3c) depicts the ex-
pected MSD power law behaviour: as the track width in-
creases αx is found to decrease monotonically to a width-
independent minimum (Fig. 3d). By contrast, log-log
MSD(y)-time (Fig. 3e) attains a slope of αy ≈ 0 for
narrow track widths in long-time limits (Fig. 3f ).
Fig. 3b also shows that the large track width values
of αr depend on polyvalency and span. Short span and
large polyvalency results in the highest αr = 1.4, whereas
the lowest αr = 1.1 is found by lowering polyvalency and
increasing the span. All of our examined BBR designs
display similar MSD trends as a function of track width,
as shown in Fig. S1.
4
FIG. 2. Sample trajectories and distributions of polyvalent
BBRs on a two-dimensional track. Left column: sample tra-
jectories. Green and red points indicate the starting and de-
tachment positions for each respective trajectory; the time
listed is the lifetime of that trajectory.
(3,3), (12,3) and
(12,8) BBRs are less likely to cross previously visited terri-
tories and exhibit more directional walks compared to (3,8)
BBRs. Right column: snapshot of ensemble BBR position
distributions, taken from Movies S1-S4. Color coding rep-
resents the number of independent BBRs at that location in
time. All BBR designs, except for (3,8), develop ring-like
distributions with decreased occupancy at the origin.
C. Detachment curves
In our simulations we do not allow for re-attachment
once all of the legs of the BBR have detached. Thus, it
is useful to investigate how track association times vary
with polyvalency and span. Fig. 4ab display the frac-
tion of BBRs remaining bound for all examined BBR
designs in the 1D and 2D track limits. We find that
(3,8) BBRs remain associated to the track for the longest
times, whereas (12,3) BBRs detach the fastest. In Fig.
4c we show that track association time increases mono-
Counts
18
17
16
15
14
13
12
11
10
12345678
Counts
25
24
22
20
19
18
16
14
13
12
10
124678
Counts
17
16
15
14
13
12
11
10
123456789
Counts
16
15
14
13
12
11
10
1234567889
(3,3)
-20
-10
0
10
20
(3,8)
-40
-20
0
20
40
(12,3)
-20
0
20
(12,8)
20
10
0
-10
-20
40
20
0
-20
-40
20
0
-20
20
10
0
-10
-20
Start
End
(3,3)
790 s
Start
End
-8
-6
-4
-2
0
2
4
Start
End
(3,8)
25000 s
-100
-80
-60
-40
-20
0
Start
End
(12,3)
1650 s
10
15
20
25
30
0
5
(12,8)
1915 s
Start
End
0246
-2
-4
-6
-8
0
-50
-100
30
25
20
15
10
05
200
150
100
50
0
y Coordinate
-50
0
50
x Coordinate
-20
-10
0
10
20
5
signs. For (12,3) ratchets, increasing track width from
one to two dimensions results in an increase of t1/2 by a
factor of 10 (Fig. 4d). A similar comparison of (12,8)
BBRs yields a factor of 60 increase in track association
time. Therefore, in the limit of large polyvalency and
short span we see less of a gain in track attachment time
by increasing the track width than for larger span.
We next compared the different designs directly by tak-
ing ratios of the t1/2 trends as shown in Fig. 4e. In the
limit of low polyvalency, increasing span from 3 to 8 has
a profound effect on increasing track association time.
Similarly, in the limit of high span, decreasing polyva-
lency from 12 to 3 profoundly increases track association
time. However, if the span is kept constant at 3, the
decrease in polyvalency from 12 to 3 has little impact
on track assocation time across all track widths. There-
fore, the improvement in track association time gained
from decreasing polyvalency is only realized for the BBR
systems with large span.
In Fig. 4f we plot αx against t1/2. For all BBR
systems αx tends to decrease with increasing t1/2 values.
D. Excess kurtosis
The ensemble-average displacement of the ratchets
away from the origin results in evolving displacement dis-
tributions, as shown in Movies S5 and S6. All of the
BBRs are initialized at the centre of the track, therefore
the initial distribution is peaked at the origin at t = 0
s. As BBRs progress along the track, their cleavage of
track sites limits options for turning back. In one dimen-
sion, as the BBRs randomly break the track symmetry
a bimodal distribution develops whose modes propagate
in opposite directions. Fig. 5a illustrates the typical
development and separation of the two modes on a nar-
row track width of 8 for (3,8) BBRs. From Movie S5,
qualitatively, one can see that the modes are both sep-
arating and dispersing with time. Fig. 5b shows the
results of computing the excess kurtosis, γ2(x), for these
(3,8) BBRs on all track widths. In all cases γ2(x) initial-
izes slightly higher than the Gaussian value of 0 as the
distribution is initially peaked sharply around the origin.
On narrow tracks γ2(x) rapidly reduces to the -2.0 limit,
indicating a distribution whose probability is located at
the edges of its domain. For wider tracks, γ2(x) reduces
to -0.25. Similar behaviour is seen for the other BBR
designs, as shown in Fig. S3.
We next look at the correponding evolution of position
distributions for the y coordinates. The y-position distri-
bution for (3,8) BBRs on a track width of 128 evolves into
a uniform distribution across the accessible domain of lat-
tice sites. To understand how this lateral shape of the
distribution changes upon interacting with the boundary
we compute γ2(y) (Fig. 5d). For wide tracks, where the
BBRs do not reach the boundaries, we see that γ2(y) ap-
proaches -0.25, the same as found for γ2(x). However, for
(3,8) BBRs on a track width of 128, which begin to ap-
FIG. 3. a) log-log MSD(~r)-time plots for (3,8) BBRs display
the general trend of decreasing slope as a function of increas-
ing track width. b) αr for (3,8) BBRs reaches a minimum at
track widths of 128-256 before subsequently increasing to a
width-invariant plateau of αr ≈ 1.1. A minimum αr at in-
termediate widths is observed for all BBR designs examined
in this work. c) MSD(x) for (3,8) BBRs scales from ballis-
tic to superdiffusive as the track width increases from one to
two dimensions. d) αx as a function of width for all BBRs
displays a monotically decreasing trend. e) MSD(y) for (3,8)
evolves to αy = 0 for all BBRs that are constrained by track
boundaries. f) For narrow tracks the BBRs move under con-
finement and display αy ≈ 0. As track width increases such
that the BBRs do not interact with the boundaries, αy takes
on the same values as αx for all BBR designs.
tonically as a function of increasing track width for (12,8)
BBRs. The detachment curves saturate and overlap for
track widths larger than 256. All BBR designs display a
similar trend, and can be viewed in Fig. S2.
For all detachment curves we define t1/2 as the time at
which 50% of the BBRs have detached from the track.
For (3,8) BBRs we observed negligible detachment on
tracks of width greater than 32, therefore we cannot re-
port t1/2 values for wider tracks. Fig. 4d depicts t1/2
as a function of track width for each BBR design. We
can see that both polyvalency and span have large ef-
fects on the observed t1/2. Across all track widths (12,3)
BBRs consistently detach faster than all other BBR de-
6
FIG. 5. a) The evolving x-position distribution for (3,8) BBRs
on a track width of 8. The numbers 1-4 represent the dis-
tribution at specific times. Two modes develop that move
symmetrically in opposing directions. b) Excess kurtosis,γ2,
for x-displacement distributions for (3,8) BBRs on all track
widths. On narrow tracks γ2(x) quickly approaches -2.0. As
track width increases the excess kurtosis settles to γ2(x) =
-0.25. The numbers 1-4 correspond to the histograms from
a). c) The evolving y-position distribution for (3,8) BBRs on
a track of width 128. d) For large track widths γ2(y) limits
to -0.25, the same as γ2(x). Under confinement γ2(y) settles
to -1.2, the value for a uniform distribution, as illustrated by
the numbers 1-4 from part c).
distributions with equal variance and equal but opposite
means, as shown in eq. 6.
P (xµ, σ) =
1
2√2πσ2
e− (x−µ)2
2σ2 +
1
2√2πσ2
e− (x+µ)2
2σ2
(6)
We computed the 2nd and 4th moments of P (x) to de-
termine the excess kurtosis (see supporting information).
We find γ2 to be given by
γ2 =
2µ4 + 12µ2σ2 + 6σ4
2(µ2 + σ2)2
− 3.
(7)
(8)
FIG. 4. a) Fraction of BBRs remaining bound in one dimen-
sion for all BBR designs examined in this work. The inset
shows the early time behaviour. b) Fraction of BBRs remain-
ing bound in two dimensions for all BBR designs examined.
For both 1D and 2D tracks (3,8) BBRs remain associated to
the track for the longest times, whereas (12,3) BBRs asso-
ciate to the track for the shortest times. c) Fraction of (12,8)
BBRs remaining bound for all track widths. The detachment
curves saturate past track widths of 256. d) t1/2 as a function
of track width for all BBR designs. e) The ratio of t1/2 values
for various BBR designs. f) αx values from Fig. 3d plotted
against their respective t1/2 values.
proach the track boundary around ∼ 1000s, we see that
γ2(y) decreases to -1.2, the value for a uniform distribu-
tion [36]. For these (3,8) BBRs γ2(x) remains constant
at -0.25 once they have reached the y-boundaries. The
shape of the x-displacement distribution is therefore not
affected by interactions with the track boundary, while
the y-displacement distribution clearly is. Similar be-
haviour is seen for the other BBRs (Fig. S4).
E. Bimodal Gaussian model
The MSD of this distribution is
To further our understanding of the time dependence
of γ2(x) we analytically derived the excess kurtosis for a
probability density function comprised of two Gaussian
MSD =(cid:10)(x − µ)2(cid:11) = µ2 + σ2.
We examine four cases in our analytical model:
a) Constant separation, Brownian dispersion :
µ = µ0; σ2 = Dt
b) Linear separation, constant dispersion :
µ = bt; σ2 = D0
c) Linear separation, Brownian dispersion :
µ = bt; σ2 = Dt
d) Linear separation & dispersion :
µ = bt; σ = D1t
For each of the above cases, eq. 8 can be expressed as
a quadratic equation with different coefficients,
MSD = a0 + a1t + a2t2.
(9)
Tuning the linear and quadratic coefficients can be used
to tailor the power-law scaling α. For example, if b = 0
and D > 0 we achieve conventional diffusion (α = 1),
whereas if D = 0 and b > 0 we get ballistic motion
(α = 2). If both b > 0 and D > 0 the type of diffusion
depends on the ratio D
b , which dictates the timescales of
interest.
In Fig. 6 we take D = D1 = b = 1, µ0 = D0 = 10,
and compute MSD (Fig. 6a) and γ2 (Fig. 6b). For
case a, we find the system exhibits subdiffusive motion
for timescales up to 103 seconds (Fig. 6a). We do not
see conventional diffusion unless we compute α at longer
times. Similarly, for case c, subdiffusive behaviour is ob-
served at short timescales followed by a crossover to bal-
listic motion at longer times.
The excess kurtosis also varies by case. For case a,
where the mean of the modes is fixed to ±µ0 with Brow-
nian dispersion, γ2 increases from -2.0 to the Gaussian
limit of 0. With cases b and c, where the modes are sep-
arating faster than they are dispersing, γ2 reduces from
0 to -2.0. Lastly for case d, where µ(t) and σ(t) are
both linearly increasing at the same rate, γ2 takes on a
constant value of -0.5.
F. Substrate digestion rates
Those who study starved random walks are often con-
cerned with the number of food items the walker con-
sumes before starvation [7]. In our model starvation can
be defined as the walker having no accessible substrate
'food' sites within its span. While we have already char-
acterized the total time associated to the tracks, here
we characterise track digestion rates and total successful
cleavages before starvation (detachment). We define the
substrate digestion rate, kd, as the average number of
cleavages observed for each BBR design per second. kd
is distinct from kef f which is known a priori and used
to simulate cleavage kinetics in the Gillespie algorithm.
In Fig. 7a we report kd for all examined BBRs across
all track widths. Not surprisingly, we find polyvalency
7
FIG. 6. Results of bimodal Gaussian model. a) In the limit of
Brownian dispersion (case a) we recover conventional diffusion
at long times. For all other cases we find the long-time limit
of α to represent ballistic motion. b) With increasing σ(t) and
constant µ (case a), γ2 reaches the Gaussian limit of γ2 = 0.
In cases b and c, which describe distributions that separate
faster than they disperse, γ2 decreases to -2.0. When σ and
µ increase at the same rate γ2 remains constant (case d).
to be the dominating factor for increasing the digestion
rate, where both (12,8) and (12,3) BBRs have the highest
kd. Interestingly, cleavage rates for the 12-legged BBRs
increase with track width to a constant value, whereas
3-legged BBRs experience a slight decrease in their cleav-
age rates as the width of the track is increased. When we
examine the average number of cleavages before detach-
ment, the inherent track association time, characterised
by t1/2, is the dominating system parameter. Fig. 7b
displays average cleavage events vs. track width, which
scales similary to t1/2 (Fig. 4d).
FIG. 7. a) Digestion rate, kd, for each BBR design across all
track widths. For all BBR designs substrate digestion rates
are found to be width-independent for tracks larger than a
width of ∼8. b) Average number of substrate sites digested
before detachment (starvation) for all BBRs across all ex-
amined track widths. For (3,8) BBRs, average cleavages are
reported for tracks only up to width 32, as most (3,8) BBRs
remained bound on larger track widths.
IV. DISCUSSION
A. Mean squared displacement
The MSD holds information on the BBR's ability to
move directionally, a key design criterion for synthetic
molecular motors. MSD scales as tα, where α charac-
terizes the type of anamolous diffusion inherent to the
system [35]. For values of α ranging from 0 < α < 1 the
motion is subdiffusive, which describes characteristic mo-
tion slower than that of conventional diffusion. A value
of α = 1 describes a system that undergoes conventional
diffusion. Systems with 1 < α < 2 are superdiffusive, a
property of systems that undergo active transport. When
α = 2 the system exhibits linear (ballistic) motion and is
ideal for molecular transport systems.
We find the long-time α values to be highly dependent
on track width, span, and polyvalency. We expected α
to maximize on narrow tracks for each BBR design, as
the effect of confinement would promote linearly directed
motion. With reference to Fig. 3bd, we find that α is
maximum for narrow tracks and persists with near bal-
listic values for tracks of width larger than 1 lattice site.
For (12,8) BBRs, αx remains constant for track widths
up to 16, whereas for (3,8) BBRs αx begins to decrease
at a track width of 4 (half the span for this system) (Fig.
3d). Therefore, both polyvalency and span play a role
in maintaining optimally ballistic motion as the effects of
confinement are relaxed.
1
At early times all examined BBRs display subdiffu-
sive behaviour at timescales proportional to
, as was
[32]. We observe initially
also reported by Olah et al.
subdiffusive behaviour for all track widths. This makes
intuitive sense as each run is initialized with one leg as-
sociated to the track. Unbound legs then need to bind
and cleave in order to translocate the global constraint,
which by design occurs on a timescale of
regardless
of the effects of confinement.
kef f
1
kef f
In Fig. 3f we report αy values from fitting to the
late-time MSD trends when the BBRs have had suffi-
cient time to reach the boundaries. As the track width
increases, the characteristic time for αy to transition to 0
increases. When αy = 0, MSD(y) takes on the variance
of a uniform distribution whose domain is defined by the
track boundaries.
The surprising result in Fig. 3b, where αr devel-
ops a minimum as a function of width, can then be ex-
plained by interactions with the boundary, which impose
a sub-diffusive characteristic on αr.
In contrast, when
the BBRs are not constrained by the track boundaries
we find that αr = αx = αy.
For effectively two-dimensional tracks, larger span and
lower polyvalency, such as the (3,8) BBRs, results in a
lower αr of 1.1. The evolution of the ensemble of (3,8)
BBRs, as shown in Movie S2, also displays a perma-
nent high occupancy around the starting position. Con-
versely, higher polyvalency and shorter span, such as the
8
(12,3) BBRs, results in a greater αr of 1.4. For the
(12,3) system, the evolution of the ensemble distribu-
tion (Movie S3) also displays low occupancy around the
starting position, leading to a ring-like structure in the
two-dimensional distribution. The emergence of the ring
structure indicates that the ensemble exhibits radially
directed motion, away from the starting position.
Increased polyvalency therefore leads to the most su-
perdiffusive walk in two dimensions. Why is this so? In
our Gillespie model, the rate of binding to substrate, kon,
is ∼400 times higher than the effective rate of cleavage,
kef f . Therefore, all unbound legs will preferentially bind
to locally available substrate sites. Each unbound leg
acquires a transition rate, kon, for each of the available
N substrate sites. For (12,3) BBRs this means that all
legs will preferentially saturate the track. The increased
number of track-associated legs means that the product
wake produced by the BBR is also denser. By contrast,
(3,8) BBRs can access larger regions of the track with
each step, and with a polyvalency of only 3 their prod-
uct wake is expected to be sparse. By this reasoning,
the (12,3) BBRs are expected to have higher αr values
than the (3,8) BBRs because it is harder for the global
constraint to change direction; multiple legs need to co-
ordinate to move the global constraint towards a new
direction, leading to a higher value of αr.
Substrate track sites that have previously been visited,
and subsequently turned to product-sites, cannot be re-
visited. However, the global constraint can still visit its
previous locations and cross over its path because the
legs can bind beyond their nearest neighbours. The more
times the global constraint revisits a location the less
likely it will be to return because the local region be-
comes further depleted of substrate. Therefore, despite
the ideal burnt-bridges behaviour of each leg, our BBRs
do not scale as a strict self-avoiding walk. There may be
merit in the application of models for weakly self-avoiding
walks to polyvalent BBRs [37]. The BBR system may
also bear relevance to foragers eating a subset of food
per site [38].
B. Track dissociation and its effect on α
The width of the track has a strong effect on both
the observed α values and the track association time, as
shown in Fig. 3 and Fig. 4. As stated αx does not be-
gin to decrease until a track width of 16 for (12,8) BBRs
(Fig. 3d). However, their track association time in-
creases dramatically from a width of 1 to a width of 16
(Fig. 4c). To understand the relationship between αx
and t1/2 we first looked to see if they are correlated. In
Fig. 4f we see that αx is relatively constant and inde-
pendent of t1/2 for these BBRs on tracks of width 1-16.
t1/2 ≈ 100 s on a track of width 1, whereas on a track of
width 16 t1/2 ≈ 2000 s. By increasing the width of the
track to twice that of the BBR span, the (12,8) system
maintains ballistic behaviour while gaining more than an
order-of-magnitude increase in track attachment time. A
further increase in track width from 16 to 4096 results in
an increase of t1/2 by a factor of 2.5, but a decrease in
αx from ∼ 2.0 to ∼ 1.3. We therefore conclude that if
one wishes to increase both directionality and track as-
sociation time, designing tracks of a width proportional
to the span of the ratchet is optimal. When one further
increases the track width, α begins to decrease.
Both decreasing polyvalency and increasing span re-
sult in increased track attachment time. Of the two
design parameters, which has the strongest impact on
maintaining track assocation? To illustrate the effects of
polyvalency on track association we can compare (3,8)
and (12,8) BBRs on tracks of width 1 and 8. On a track
width of 1, we find t1/2 to be 240 and 110 seconds for (3,8)
and (12,8) BBRs, respectively (Fig. 4d). However, the
effect of polyvalency on track association highly depends
on the span, which we map out in Fig. 4e. For example,
increasing polyvalency of span-3 BBRs results in at most
a factor of 2 increase in t1/2, even on two-dimensional
tracks. However, for span-8 BBRs, t1/2 increases by a
factor of 2 for one-dimensional tracks, but by a factor of
10 for two-dimensional tracks when the number of legs
is increased from 3 to 12. Altering span also shows sim-
ilar trends where the gain in t1/2 highly depends on the
polyvalency.
It may seem counterintuitive that an increase in poly-
valency leads to a decrease in t1/2 given previous work
with molecular spiders [16, 31]. However, in molecular
spider systems the walkers have a rate of product bind-
ing, konP , typically taken to be the same as substrate
binding, konS [30–32]. Therefore, when the walker di-
gests all local substrate sites it can search through local
product sites for areas of fresh substrate. In such a sys-
tem, increased polyvalency means more options for prod-
uct site coupling and subsequently decreased probability
of detachment. By contrast, our system is a ideal BBR
where konP = 0, and we find increased polyvalency leads
to a decrease in track attachment time.
C. Excess Kurtosis
The shape of the BBR position distribution has a time
dependence. Kurtosis is a convenient measure to com-
pare BBR position distributions across our parameter
space of width, polyvalency, and span. On narrow tracks
the position distribution for (3,8) BBRs immediately de-
velops into two modes that move in opposite directions.
This is reflected as a monotonic decrease in γ2(x) to -2.0
(Fig. 5b), which indicates a distribution with proba-
bility isolated to the edges of the domain. The splitting
of the position distribution into two oppositely moving
modes is consistent with the ballistic behaviour charac-
terized by α.
Having seen that the distributions formed two oppo-
sitely moving modes inspired us to analytically derive
kurtosis for a PDF described by two Gaussians, as shown
9
in eq. 6. We analytically derived γ2 (eq. 7) in one dimen-
sion and explored the effects of dispersion and mode sep-
aration, as shown in Fig. 6. Increasing dispersion, while
maintaining a constant separation of modes, leads to γ2
monotonically increasing towards the Gaussian value of
0, indicating that the two independent modes are com-
pletely overlapping. Conversely, a linear increase in mode
separation with constant dispersion results in γ2 mono-
tonically decreasing to -2.0. Furthermore, dispersion and
mode separation have compensatory effects on γ2, such
that if they are increasing at similar rates γ2 remains
constant. From our phenomenological Gaussian model
we are able to reproduce all of the BBR γ2 results by
modifying σ and µ.
Our analytical model offers insights into the BBR be-
haviour and allows us to understand the effects of poly-
valency, span and width on the shape of the position dis-
tributions. For example, γ2(x) for (3,8) BBRs on wide
tracks reaches a time-invariant value of -0.25 (Fig. 5b).
From our bimodal Gaussian model, this suggests that
the dispersion and separation of the modes are equal.
We verify this for the (3,8) system where we compute
σ(t)/µ(t) (Fig. S5). The ratio of σ(t)/µ(t) is known as
the coefficient of variation. In our analytical system the
coefficient of variation is equal to D
b where D and b are
equivalent to diffusion and drift coefficients, respectively.
The ratio D
b is a ratio of diffusion and mobility, and has
been used to characterize the ability of Brownian ratchets
to achieve directional motion under external fields [39].
On wide tracks, for all BBR designs, γ2(y) takes on
the same values as γ2(x) (Figs. S3,S4). As the BBRs
are constrained by the boundaries, the position distribu-
tions evolve into uniform distributions across the width
(γ2(y) = −1.2), consistent with our finding that the vari-
ance given by MSD(y) approaches that of a uniform dis-
tribution.
D. Substrate digestion rates
Across all track widths we find that BBRs with larger
span and polyvalency have the highest substrate diges-
tion rate (kd), as shown in Fig. 7a. The ratio of the
binding and effective cleavage rates used in the Gillespie
model, kon and kef f , respectively, is ∼400. This means
that an unbound leg that can access a fresh substrate site
is going to be 400 times more likely to bind to any one site
than a bound leg is to cleave and release. Thus, all legs
are likely to be bound to available substrate sites. (12,8)
BBRs have access to more local substrate than the other
BBR designs given that they have the largest polyvalency
and the longest reach (span), therefore (12,8) BBRs are
expected to cleave the most per unit time. (12,3) BBRs
are 'second best' to (12,8) BBRs with regards to the sub-
strate digestion rate. All 12 legs can saturate to the track,
but this design suffers from a shorter span leading to an
inability of (12,3) BBRs to reach distant patches of fresh
substrate, as compared to the (12,8) system.
It was surprising to see that (3,3) and (3,8) BBRs ex-
perienced a slight decrease in average substrate digestion
rates as a function of increasing track width (Fig. 7a).
from 0.18 s−1 in one dimension,
The decline is slight:
to 0.16 s−1 in two dimensions for (3,3) BBRs. As the
width is increased one would naively think that more
substrate should be available to the BBR for any given
combination of bound legs, therefore resulting in an in-
crease digestion rate. This was not found for three-legged
BBRs. Furthermore, for any given track width, the av-
erage digestion rates for (3,3) and (3,8) BBRs are within
5%, suggesting that in the limit of low polyvalency, span
plays no significant role in altering the substrate diges-
tion rate.
Our hypothesis is that in one dimension, when the
span-3 walkers move into their product wake, they
quickly detach as there is little opportunity to turn
around towards fresh substrate. However, as the track
width is increased the walkers have more opportunity to
rescue themselves from a substrate-barren environment.
We speculate that on average, the (3,8) and (3,3) walkers
experience more substrate-barren terrain in wider track
widths, which leads to a lower average substrate diges-
tion rate as they spend more time rescuing themselves
from locally depleted regions.
The BBRs studied in this work can be considered as
polyvalent depletion-controlled foragers [7, 40]. In Fig.
7b we plot the average number of substrate sites digested
by each BBR on each track width. As a function of track
width, substrate digestion per lifetime scales similarly as
t1/2 (Fig. 4d). Despite (3,8) BBRs having the lowest
kd across all track widths, their greater track association
time leads to them having more time to digest substrate,
resulting in the most substrate cleaved prior to detach-
ment.
V. CONCLUSIONS
10
The design and implementation of synthetic machinery
has shown great promise towards the control of motion
at the nanoscale. In particular, synthetic analogues of bi-
ological molecular motors that implement a BBR mech-
anism have made great progress. Our goal in this work
was to explore the effects of confinement on BBR perfor-
mance and to provide design insights for de novo BBR
motors. Our results offer guidelines for researchers to
follow when thinking about optimizing particular BBR
characteristics. To fabricate a superdiffusive BBR in two
dimensions, one should increase polyvalency and decrease
span, as has been done in some systems [11, 25].
In-
creasing span and decreasing polyvalency, in contrast,
results in large increases to track attachment time but
decreased directionality. Furthermore, we found that
narrow tracks result in ballistic dynamics, as well as
an order-of-magnitude increase in track attachment time
compared to a one-dimensional track. Lastly, we found
that increasing polyvalency results in an increased rate
of substrate digestion, however, the total average track
association time is the dominant factor that dictates to-
tal cleavage events before detachment. Through explor-
ing the dimensionality-dependent crossover in motility of
polyvalent BBRs, we have found these systems to exhibit
rich dynamics. We hope these results provide useful in-
sight towards the design of de novo BBR systems.
VI. ACKNOWLEDGEMENTS
This work was funded by the Natural Sciences and En-
gineering Research Council of Canada (NSERC), through
a Discovery Grant to NRF. Computational resources
were provided by Compute Canada.
[1] K. Svobody, F. C. Schmidt, J. B. Schnapp, and M. S.
Block, "Direct Observation of kinesin stepping by optical
trapping interferometry," Nature, vol. 365, pp. 721–727,
1993.
[2] H. Nobutaka, "Kinesin and dynein superfamily proteins
and the mechanism of organelle transport," Science,
vol. 279, no. 5350, pp. 519–526, 1998.
[3] D. A. Mehta, S. R. Rock, M. Rief, A. J. Spudich, S. M.
Mooseker, and E. R. Cheney, "Myosin-V is a processive
actin-based motor," Nature, vol. 400, pp. 590–593, 1999.
[4] M. Schliwa and G. Woehlke, "Molecular motors," Nature,
vol. 422, pp. 759–65, 2003.
[5] L. Bruno, V. Levi, M. Brunstein, and M. A. Despos-
ito, "Transition to superdiffusive behavior in intracellular
actin-based transport mediated by molecular motors,"
Phys. Rev. E, vol. 80, no. 1, p. 011912, 2009.
[6] T. Antal and P. L. Krapivsky, "'Burnt-bridge' mechanism
of molecular motor motion," Phys. Rev. E, vol. 72, no. 4,
p. 046104, 2005.
[7] O. B´enichou and S. Redner, "Depletion-controlled star-
vation of a diffusing forager," Phys. Rev. Lett., vol. 113,
no. 23, p. 238101, 2014.
[8] S. Saffarian, I. E. Collier, B. L. Marmer, E. L. Elson,
and G. Goldberg, "Interstitial collagenase is a brown-
ian ratchet driven by proteolysis of collagen," Science,
vol. 306, pp. 108–111, 2004.
[9] S. K. Sarkar, B. Marmer, G. Goldberg, and K. C. Neu-
man, "Single-molecule tracking of collagenase on native
type I collagen fibrils reveals degradation mechanism,"
Curr. Biol., vol. 22, no. 12, pp. 1047–1056, 2012.
[10] I. E. Collier, W. Legant, B. Marmer, O. Lubman, S. Saf-
farian, T. Wakatsuki, E. Elson, and G. I. Goldberg, "Dif-
fusion of MMPs on the surface of collagen fibrils: the
mobile cell surface-collagen substratum interface.," PLoS
One, vol. 6, no. 9, p. e24029, 2011.
[11] A. G. Vecchiarelli, K. C. Neuman, and K. Mizuuchi, "A
propagating ATPase gradient drives transport of surface-
confined cellular cargo.," Proc. Natl. Acad. Sci. U. S. A.,
vol. 111, no. 13, pp. 4880–4885, 2014.
[12] L. Hu, A. G. Vecchiarelli, K. Mizuuchi, K. C. Neuman,
and J. Liu, "Brownian Ratchet Mechanism for Faithful
Segregation of Low-Copy-Number Plasmids," Biophys.
J., vol. 112, no. 7, pp. 1489–1502, 2017.
[13] K. Gerdes, M. Howard, and F. Szardenings, "Push-
ing and pulling in prokaryotic DNA segregation.," Cell,
vol. 141, no. 6, pp. 927–942, 2010.
[14] S. M. Block, L. S. Goldstein, and B. J. Schnapp, "Bead
movement by single kinesin molecules studied with op-
tical tweezers," Nature, vol. 348, no. 6299, pp. 348–352,
1990.
[15] P. Yin, H. Yan, X. G. Daniell, A. J. Turberfield, and
J. H. Reif, "A unidirectional DNA walker that moves
autonomously along a track," Angew. Chemie, vol. 43,
pp. 4906–4911, 2004.
[16] R. Pei, S. K. Taylor, D. Stefanovic, S. Rudchenko, T. E.
Mitchell, and M. N. Stojanovic, "Behavior of polycat-
alytic assemblies in a substrate-displaying matrix," J.
Am. Chem. Soc., vol. 128, no. 24, pp. 12693–12699, 2006.
[17] J. Bath and A. J. Turberfield, "DNA nanomachines.,"
Nat. Nano., vol. 2, no. 5, pp. 275–84, 2007.
[18] E. H. C. Bromley, N. J. Kuwada, M. J. Zuckermann,
R. Donadini, L. Samii, G. A. Blab, G. J. Gemmen, B. J.
Lopez, P. M. G. Curmi, N. R. Forde, D. N. Woolfson, and
H. Linke, "The tumbleweed: Towards a synthetic protein
motor," HFSP J., vol. 3, no. 3, pp. 204–212, 2009.
[19] T. Omabegho, R. Sha, and N. C. Seeman, "A Bipedal
DNA Brownian Motor with Coordinated Legs," Science,
vol. 324, pp. 67–71, 2009.
[20] K. Lund, A. J. Manzo, N. Dabby, N. Michelotti,
A. Johnson-Buck, J. Nangreave, S. Taylor, R. Pei, M. N.
Stojanovic, N. G. Walter, E. Winfree, and H. Yan,
"Molecular robots guided by prescriptive landscapes.,"
Nature, vol. 465, no. 7295, pp. 206–210, 2010.
[21] J. Cheng, S. Sreelatha, R. Hou, A. Efremov, R. Liu,
J. R. C. van der Maarel, and Z. Wang, "Bipedal
Nanowalker by Pure Physical Mechanisms," Phys. Rev.
Lett., vol. 109, p. 238104, 2012.
[22] T.-G. Cha, J. Pan, H. Chen, J. Salgado, X. Li, C. Mao,
and J. H. Choi, "A synthetic DNA motor that transports
nanoparticles along carbon nanotubes," Nat. Nanotech-
nol., vol. 9, no. 1, pp. 39–43, 2014.
[23] C. S. Niman, M. J. Zuckermann, M. Balaz, J. O. Tegen-
feldt, P. M. G. Curmi, N. R. Forde, and H. Linke, "Flu-
idic switching in nanochannels for the control of Inch-
worm: a synthetic biomolecular motor with a power
stroke," Nanoscale, vol. 6, pp. 15008–15019, 2014.
[24] S. Kovacic, L. Samii, P. M. Curmi, H. Linke, M. J.
Zuckermann, and N. R. Forde, "Design and construc-
tion of the Lawnmower, an artificial burnt-bridges mo-
tor," Trans. NanoBioscience, vol. 14, no. 3, pp. 305–312,
2015.
11
[25] K. Yehl, A. Mugler, S. Vivek, Y. Liu, Y. Zhang, M. Fan,
E. R. Weeks, and K. Salaita, "High-speed DNA-based
rolling motors powered by RNase H," Nat. Nano., vol. 11,
no. 2, pp. 184–190, 2015.
[26] M. Liu, J. Cheng, S. R. Tee, S. Sreelatha,
I. Y.
Loh, and Z. Wang, "Biomimetic Autonomous Enzymatic
Nanowalker of High Fuel Efficiency," ACS Nano, vol. 10,
no. 6, pp. 5882–5890, 2016.
[27] J. Li, A. Johnson-Buck, Y. R. Yang, W. M. Shih, H. Yan,
and N. G. Walter, "Exploring the speed limit of toehold
exchange with a cartwheeling DNA acrobat," Nat. Nan-
otechnol., p. 1, 2018.
[28] T. Antal, P. L. Krapivsky, and K. Mallick, "Molecular
Spiders in One Dimension.," J. Stat. Mech., vol. 2007,
p. P08027, 2007.
[29] B. Shtylla and J. P. Keener, "Mathematical modeling of
bacterial track-altering motors: Track cleaving through
burnt-bridge ratchets," Phys. Rev. E - Stat. Nonlinear,
Soft Matter Phys., vol. 91, no. 4, pp. 1–18, 2015.
[30] L. Samii, H. Linke, M. J. Zuckermann, and N. R.
Forde, "Biased motion and molecular motor properties
of bipedal spiders," Phys Rev E, vol. 81, no. 2, p. 021106,
2010.
[31] L. Samii, G. A. Blab, E. H. C. Bromley, H. Linke,
P. M. G. Curmi, M. J. Zuckermann, and N. R. Forde,
"Time-dependent motor properties of multipedal molec-
ular spiders," Phys Rev E, vol. 84, no. 3, p. 031111, 2011.
[32] M. J. Olah and D. Stefanovic, "Superdiffusive transport
by multivalent molecular walkers moving under load,"
Phys. Rev. E, vol. 87, no. 6, p. 062713, 2013.
[33] L. Hu, A. G. Vecchiarelli, K. Mizuuchi, K. C. Neuman,
and J. Liu, "Directed and persistent movement arises
from mechanochemistry of the ParA / ParB system,"
PNAS, vol. 112, no. 51, pp. E7055–E7064, 2015.
[34] D. Gillespie, "Exact stochastic simulation of coupled
chemical reactions," J. Phys. Chem., vol. 81, no. 25,
pp. 2340–2361, 1977.
[35] S. Havlin and D. Ben-Avraham, "Diffusion in disordered
media," Adv. Phys., vol. 51, no. 1, pp. 187–292, 2002.
[36] L. T. DeCarlo, "On the Meaning and Use of Kurtosis,"
Psychol. Methods, vol. 2, no. 3, pp. 292–307, 1997.
[37] N. Madras and G. Slade, "The Self-Avoiding Walk,"
Birkhauser, pp. 365–368, 1996.
[38] O. B´enichou, U. Bhat, P. L. Krapivsky, and S. Red-
ner, "Optimally frugal foraging," Phys. Rev. E, vol. 97,
p. 022110, 2018.
[39] A. van Oudenaarden, "Brownian Ratchets: Molecular
Separations in Lipid Bilayers Supported on Patterned Ar-
rays," Science, vol. 285, no. 5430, pp. 1046–1048, 1999.
[40] O. B´enichou, M. Chupeau, and S. Redner, "Role of de-
pletion on the dynamics of a diffusing forager," J. Phys.
A: Math. Theor., vol. 49, no. 39, p. 394003, 2016.
Dimensionality-dependent crossover in motility of polyvalent
burnt-bridges ratchets
Supplementary material
June 22, 2018
Chapin S. Korosec ([email protected])
Martin J. Zuckermann ([email protected])
Nancy R. Forde ([email protected])
arXiv:1806.08066v1 [physics.bio-ph] 21 Jun 2018
1 Kurtosis of a mixture of two Gaussian distributions
On narrow tracks the BBR displacement distributions produce two modes that travel in the ±x directions. To
understand how the kurtosis is expected to behave we derive γ2 for a probability density function described by two
Gaussian modes centred at µ1 and µ2 with standard deviations σ1 and σ2,
2
P (x) from eq. 1 can be used to compute the kth moment about the mean according to
e− (x−µ1 )2
2σ2
1 +
e− (x−µ2)2
2σ2
2
.
1
2p2πσ2
2
P (xµ1, µ2, σ1, σ2) =
1
1
2p2πσ2
mk =Z ∞
−∞
(x − µ)kP (x)dx.
(1)
(2)
Kurtosis is defined as the standardized 4th moment. Since a Gaussian distribution always has a kurtosis of 3, it is
convenient to define the excess kurtosis,γ2, as
(3)
(4)
(5)
(6)
2(µ4
1 + 6µ2
γ2 =
1σ2
(µ2
1 + 3σ4
1 + σ2
1 + µ4
1 + µ2
2 + 6µ2
2 + σ2
2)2
2σ2
2 + 3σ4
2)
− 3.
(µ4
1 + 6µ2
1σ2
1 + 3σ4
1 + µ4
2 + 6µ2
2σ2
2 + 3σ4
2),
(µ2
1 + σ2
1 + µ2
2 + σ2
2).
1 2
1 2
m4 =
m2 =
γ2 =
E(cid:2)(x − µ)4(cid:3)
(E [(x − µ)2])2 − 3 =
m4
(m2)2 − 3.
Using eq. 2 we find the fourth and second moments to be
Substituting eq. 4 and eq. 5 into eq. 3 we find the excess kurtosis of P (x) to be given by
In our system there is symmetry in the two modes about x = 0. Therefore, we make the further assumption
that µ2 = −µ1, so that µ2
2 = µ2
1 = µ2
0 and let σ1 = σ2 = σ0. γ2 can then be written as
γ2 =
(2µ4
0 + 12µ2
0σ2
0 + σ2
0 + 6σ4
0)
0)2
2(µ2
− 3.
(7)
2 Supplementary figures
3
Figure S1: Each panel shows the evolution of MSD(~r) for a given BBR design. MSD(~r) for each of the BBR designs
displays a similar trend with track width as discussed in the main text for (3,8) BBRs. The resulting αr values from
each MSD(~r) curve are plotted in Fig. 3b.
4
Figure S2: All detachment curves for each BBR design as a function of track width. (3,8) BBRs display negligible
detachment for track widths larger than 32. All BBRs display a similar trend of longer track association as a function
of increasing track width.
5
Figure S3: Excess kurtosis γ2(x), for all BBR designs across all track widths. a) (3,3) γ2(x) show similar trends to
that of (3,8) BBRs, however, with their shorter track association times the curves are more noisy. b) γ2(x) for (3,8)
BBRs, as shown and disucssed in the main text. c) (12,3) BBRs on narrow tracks display a decrease in γ2(x) to -2.0.
However, on wider tracks and later times the BBRs experience an increase in γ2(x), indicating that over long times
the dispersion increases at a greater rate than the modes separate. d) (12,8) BBRs display a similar trend in γ2(x)
as (12,3) BBRs, where γ2(x) increases with time on wide tracks.
6
Figure S4: Excess kurtosis γ2(y), for all BBR designs across tracks of width 16-4096. For (3,3), (12,3), and (12,8)
BBRs on intermediate tracks, γ2(y) decreases below -1.2 to a minimum before increasing to the -1.2 uniform distri-
bution limit. (12,8) BBRs on wide tracks display an appreciable increase in γ2(y) over long times. For these BBRs,
upon interacting with the confining walls γ2(y) turns over to a decreasing trend, as shown by the width 128 BBRs
((12,8), purple line). The γ2(y) trends for (3,8) BBRs are included and discussed in the main text.
7
Figure S5: µ(t) and σ(t) for BBRs with 3 legs, span 8 on a track width of 128. Left: µx(t) and µy(t) overlap until
the ensemble of motors reaches the boundary at ∼ 1000s, at which point µy(t) curves over to the expected value of
32 for a uniform distribution with boundaries at y = 0, 64. Right: The ensemble average BBRs reach the boundary
at ∼ 1000s, indicated by the red line. σx(t)
µy(t) , the coefficients of variation, indicate that the dispersion
in y-positions decreases substantially more than the mean y position upon interaction with the boundary. σx(t)
µx(t)
µx(t) and σy(t)
remains relatively constant following the onset of BBR interaction with the boundaries.
3 Supplementary movies
Movie S1: The ensemble position evolution for (3,3) BBRs on an effectively two-dimensional track. The ensemble of
BBRs can be seen to develop a low occupancy around the centre (starting point) of the track.
Movie S2: The ensemble position evolution for (3,8) BBRs on an effectively two-dimensional track. The ensem-
ble of BBRs can be seen to maintain a high occupancy around the centre (starting point) of the track.
Movie S3: The ensemble position evolution for (12,3) BBRs on an effectively two-dimensional track. The ensemble
of BBRs can be seen to develop a low occupancy around the centre (starting point) of the track.
8
Movie S4: The ensemble position evolution for (12,8) BBRs on an effectively two-dimensional track. The ensemble
of BBRs can be seen to develop a low occupancy around the centre (starting point) of the track.
Movie S5: The ensemble position evolution for (3,8) BBRs on a narrow track of width 8. Two modes can be
seen to develop and move in opposite directions.
Movie S6: The ensemble position evolution for (12,8) BBRs on a narrow track of width 8. Two modes can be
seen to develop and move in opposite directions.
|
1303.7292 | 1 | 1303 | 2013-03-29T04:06:49 | Multi-functional foot use during running in the zebra-tailed lizard (Callisaurus draconoides) | [
"physics.bio-ph",
"cond-mat.soft",
"physics.flu-dyn",
"q-bio.QM"
] | A diversity of animals that run on solid, level, flat, non-slip surfaces appear to bounce on their legs; elastic elements in the limbs can store and return energy during each step. The mechanics and energetics of running in natural terrain, particularly on surfaces that can yield and flow under stress, is less understood. The zebra-tailed lizard (Callisaurus draconoides), a small desert generalist with a large, elongate, tendinous hind foot, runs rapidly across a variety of natural substrates. We use high speed video to obtain detailed three-dimensional running kinematics on solid and granular surfaces to reveal how leg, foot, and substrate mechanics contribute to its high locomotor performance. Running at ~10 body length/s (~1 m/s), the center of mass oscillates like a spring-mass system on both substrates, with only 15% reduction in stride length on the granular surface. On the solid surface, a strut-spring model of the hind limb reveals that the hind foot saves about 40% of the mechanical work needed per step, significant for the lizard's small size. On the granular surface, a penetration force model and hypothesized subsurface foot rotation indicates that the hind foot paddles through fluidized granular medium, and that the energy lost per step during irreversible deformation of the substrate does not differ from the reduction in the mechanical energy of the center of mass. The upper hind leg muscles must perform three times as much mechanical work on the granular surface as on the solid surface to compensate for the greater energy lost within the foot and to the substrate. | physics.bio-ph | physics | Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Multi-functional foot use during running in the zebra-tailed
lizard (Callisaurus draconoides)
Chen Li1, S. Tonia Hsieh2, and Daniel I. Goldman1*
1School of Physics, Georgia Institute of Technology, Atlanta, GA 30332, USA and
2Department of Biology, Temple University, Philadelphia, PA 19122, USA
*Author for correspondence ([email protected])
Summary
A diversity of animals that run on solid, level, flat, non-slip surfaces appear to bounce on
their legs; elastic elements in the limbs can store and return energy during each step. The
mechanics and energetics of running in natural terrain, particularly on surfaces that can
yield and flow under stress, is less understood. The zebra-tailed lizard (Callisaurus
draconoides), a small desert generalist with a large, elongate, tendinous hind foot, runs
rapidly across a variety of natural substrates. We use high speed video to obtain detailed
three-dimensional running kinematics on solid and granular surfaces to reveal how leg, foot,
and substrate mechanics contribute to its high locomotor performance. Running at ~ 10
body length/s (~ 1 m/s), the center of mass oscillates like a spring-mass system on both
substrates, with only 15% reduction in stride length on the granular surface. On the solid
surface, a strut-spring model of the hind limb reveals that the hind foot saves about 40% of
the mechanical work needed per step, significant for the lizard’s small size. On the granular
surface, a penetration force model and hypothesized subsurface foot rotation indicates that
the hind foot paddles through fluidized granular medium, and that the energy lost per step
during irreversible deformation of the substrate does not differ from the reduction in the
mechanical energy of the center of mass. The upper hind leg muscles must perform three
times as much mechanical work on the granular surface as on the solid surface to
compensate for the greater energy lost within the foot and to the substrate.
Key words: terrestrial locomotion, mechanics, energetics, kinematics, spring-mass system, elastic
energy savings, dissipation, granular media
Running title: Substrate effects on foot use in lizards
1
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Introduction
Rapid locomotion like running and hopping can be modeled as a spring-mass system bouncing in
the sagittal plane (i.e., the Spring-Loaded Inverted Pendulum model, SLIP) (Blickhan, 1989).
This has been demonstrated in a variety of animals (Blickhan and Full, 1993; Holmes et al., 2006)
in the laboratory on rigid, level, flat, non-slip surfaces (hereafter referred to as “solid surfaces”)
such as running tracks and treadmills (Dickinson et al., 2000). In the SLIP model, the animal
body (represented by the center of mass, CoM) bounces on a single leg (represented by a spring)
like a pogo stick, and exerts point contact on the solid ground. The leg spring compresses during
the first half of stance, and then recoils during the second half of stance. Through this process, the
mechanical (i.e., kinetic plus gravitational potential) energy of the CoM is exchanged with elastic
energy stored in the compressed leg spring, reducing energy use during each step. For animals
like insects (e.g., Schmitt et al., 2002) and reptiles (e.g., Chen et al., 2006) that run with a
sprawled limb posture, the CoM also oscillates substantially in the horizontal plane in a similar
fashion, which can also be modeled as a spring-mass system bouncing in the horizontal plane (i.e.,
the Lateral Leg Spring model, LLS) (Schmitt et al., 2002). Both the SLIP and the LLS models
predict that the mechanical energy of the CoM is lowest at mid-stance and highest during aerial
phase.
In these models, the spring-mass system and the interaction with the solid ground are perfectly
elastic and do not dissipate energy; thus no net work is performed. However, as animals move
across natural surfaces, energy is dissipated both within their body and limbs (Fung, 1993) and to
the environment (Dickinson et al., 2000). Therefore, mechanisms to reduce energy loss during
locomotion can be important. The limbs of many organisms possess elastic elements such as
tendons and ligaments that can function as springs to store and return energy during rapid
locomotion like running and hopping to decrease energetic cost (Alexander, 2003). Most notable
for this function are the ankle extensor tendons in the lower hind leg and the digital flexor
tendons and ligaments in the lower fore leg (Alexander, 2003). Furthermore, different limb-
ground interaction strategies may be utilized depending on the dissipative properties of the
substrate.
Laboratory experiments have begun to reveal mechanisms of organisms running on non-solid
surbstrates, such as elastic (Ferris et al., 1998; Spence et al., 2010), damped (Moritz and Farley,
2003), inclined (Roberts et al., 1997), or uneven (Daley and Biewener, 2006; Sponberg and Full,
2008) surfaces; surfaces with few footholds (Spagna et al., 2007); and the surface of water
(Glasheen and McMahon 1996a; Hsieh, 2003). While spring-mass-like CoM motion was
2
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
observed only in some of these studies (Ferris et al., 1998; Moritz and Farley, 2003; Spence et al.,
2010), a common finding is that on non-solid surfaces limbs do not necessarily behave like
springs to save energy. In addition, these studies suggest that both the active control of body and
limb movement through the nervous system, and the passive mechanical responses of
viscoelastic limbs and feet with the environment, play important roles in the control of rapid
terrestrial locomotion (for reviews, see Full and Koditschek, 1999; Dickinson et al., 2000).
Many substrates found in nature, such as sand, gravel, rubble, dirt, soil, mud, and debris, can
yield and flow under stress during locomotion and experience solid-fluid transitions, through
which energy may be dissipated via plastic deformation. Understanding locomotion on these
substrates is challenging in part because, unlike for flying and swimming where the fluid flows
and forces can in principle be determined by solving the Navier-Stokes equations in the presence
of moving boundary conditions (Vogel, 1996), no comprehensive force models yet exist for
terrestrial substrates that yield and flow (hereafter referred to as “flowing substrates”).
Granular materials (Nedderman, 1992) like desert sand which are composed of similarly sized
particles provide a good model substrate for studying locomotion on flowing substrates.
Compared to other flowing substrates, granular materials are relatively simple and the intrusion
forces within them can be modeled empirically (Hill et al., 2005). Their mechanical properties
can also be precisely and repeatedly controlled using a fluidized bed (Li et al., 2009). In addition,
locomotion on granular surfaces is directly relevant for many desert-dwelling reptiles and
arthropods such as lizards, snakes, and insects (Mosauer, 1932; Crawford, 1981). Recent
advances in the understanding of force and flow laws in granular materials subject to localized
intrusion (Hill et al., 2005; Katsuragi and Durian, 2007; Gravish et al., 2010; Ding et al., 2011a)
begin to provide insight into the mechanics of locomotion on (and within) granular substrates (Li
et al., 2009; Maladen et al., 2009; Mazouchova et al., 2010; Li et al., 2010b; Maladen et al., 2011;
Ding et al., 2011b; Li et al., in press).
The zebra-tailed lizard (Callisaurus draconoides, SVL ~ 10 cm, mass ~ 10 g, Fig. 1A) is an
excellent model organism for studying running on natural surfaces, because of its high locomotor
performance over diverse terrain. As a desert generalist, this lizard lives in a range of desert
habitats including flat land, washes, and sand dunes (Vitt and Ohmart, 1977; Korff and McHenry,
2011), and encounters a large variety of substrates ranging from rocks, gravel, closely-packed
coarse sand, and loosely-packed fine sand (Karasov and Anderson, 1998; Korff and McHenry,
2011). The zebra-tailed lizard is the fastest-running species among desert lizards of similar size
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
3
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
(Irschick and Jayne, 1999a), and has been observed to run at up to 4 m/s (50 bl/s) both on solid
(e.g., treadmill) (Irschick and Jayne, 1999a) and on granular (e.g., sand dunes) (Irschick and
Jayne, 1999b) surfaces. Its maximal acceleration and running speed also did not differ
significantly when substrate changes from coarse wash sand to fine dune sand, whose yield
strengths differ by a factor of three (Korff and McHenry, 2011).
Of particular interest is whether and how the zebra-tailed lizard’s large, elongate hind foot
contributes to its high locomotor capacity. In addition to a slim body, a long tapering tail, and
slender legs (Fig. 1A), the zebra-tailed lizard has an extremely large, elongate hind foot, the
largest (40% SVL) among lizards of similar size (Irschick and Jayne, 1999a). Its hind foot is
substantially larger than the fore foot (area = 1 cm2 vs. = 0.3 cm2) and likely plays a dominant role
for locomotion (Mosauer, 1932). Recent studies in insects, spiders, and geckos (Jindrich and Full,
1999; Antumn et al., 2000; Dudek and Full, 2006; Spagna et al., 2007) suggested that animals can
rely on appropriate morphology and material properties of their bodies and limbs to accommodate
variable, uncertain conditions during locomotion. Despite suggestions that the large foot area
(Mosauer, 1932) and increased stride length via elongate toes may confer locomotor advantages
(Irschick and Jayne, 1999a), the mechanisms of how the hind foot contributes to the zebra-tailed
lizard’s high running capacity remain unknown.
In this paper, we study the mechanics and mechanical energetics of the zebra-tailed lizard running
on two well-defined model surfaces, a solid surface and a granular surface. These two surfaces lie
on opposite ends of the spectrum of substrates that the zebra-tailed lizard encounters in its natural
environment, and present distinct conditions for locomotion. We investigate whether the lizard’s
center of mass bounces like a spring-mass system during running on both solid and granular
surfaces. We combine measurements of three-dimensional kinematics of the lizard’s body, hind
limb, and hind foot, dissection and resilience measurements of the hind limb, and modeling of
foot-ground interactions on both substrates, and demonstrate that the lizard’s large, elongate hind
foot serves different functions during running on solid and granular surfaces. We find that on the
solid surface, the hind foot functions as an energy-saving spring; on the granular surface, it
functions as a dissipative, force-generating paddle to generate sufficient lift during each step. The
larger energy dissipation to the substrate and within the foot during running on the granular
surface must be compensated for by greater mechanical work done by the upper hind leg muscles.
Materials and methods
4
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Animals
Seven adult zebra-tailed lizards (Callisaurus draconoides) were collected from the Mojave Desert,
AZ, USA in 08/2007 (Permit SP591773) for three-dimensional kinematics experiments. Table 1
shows the morphological measurements for these seven animals. Eleven additional adult animals
were collected from the Mojave Desert, CA, USA in 09/2009 (Permit SC 10901) for hind limb
resilience measurements. Two preserved specimens were used for dissection. The animals were
housed in the Physiological Research Laboratory animal facility of The Georgia Institute of
Technology. Each animal was housed individually in an aquarium filled with sand, and fed
crickets and mealworms dusted with vitamin and calcium supplement two to three times a week.
The ambient temperature was maintained at 28°C during the day and 24°C during the night. Full-
spectrum fluorescent bulbs high in UVB were set to a 12 hour/12 hour light/dark schedule.
Ceramic heating elements were provided 24 hours a day to allow the animals to thermo-regulate
at preferred body temperature. All experimental procedures were conducted in accordance with
The Georgia Institute of Technology IACUC protocols.
Surface treatments
A wood board (120 × 23 × 1 cm3) bonded with sandpaper (grit size ~ 0.1 mm) for enhanced
traction was used as the solid surface. Glass particles (diameter = 0.27 ± 0.04 mm mean ± 1
standard deviation, density = 2.5 × 103 kg/m3, Jaygo Incorporated, Union, NJ, USA) were used as
the granular substrate, which are approximately spherical and of similar size to typical desert sand
(Dickinson and Ward, 1994). Before each trial, a custom-made fluidized bed trackway (200 cm
long, 50 cm wide) prepared the granular substrate (12 cm deep) into a loosely packed state
(volume fraction= 0.58) for repeatable yield strength (for experimental details of the fluidized
bed trackway, see Li et al., 2009).
Three-dimensional kinematics
We used high speed video to obtain three-dimensional kinematics as the lizard ran across the
prepared surfaces (Fig. 1B). Before each session, high-contrast markers (Wite-Out, Garden Grove,
CA, USA) were painted on each animal for digitizing at nine joints along the midline of the trunk
and the right hind limb (Fig. 1A,B): neck (N), center of mass (CoM), pelvis (P), hip (H), knee (K),
ankle (A); and the metatarsal-phalangeal joint (MP), distal end of the proximal phalanx (PP), and
digit tip (T) of the fourth toe. The approximate longitudinal location of the CoM in resting
position was determined by tying a thread around the body of an anesthetized lizard and
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
5
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
repositioning the thread until the body balanced horizontally. Before each trial, the surface was
prepared (for the granular surface treatment only), and calibration images were taken of a custom-
made 39-point calibration object (composed of LEGO, Billund, Denmark). The animal was then
induced to run across the field of view by a slight tap or pinch on the tail. Two synchronized AOS
high speed cameras (AOS Technologies, Baden Daettwil, Switzerland) captured simultaneous
dorsal and lateral views at 500 frame/s (shutter time = 300 s). The ambient temperature was
maintained at 35°C during the test. Animals were allowed to rest at least five minutes between
trials and at least two days between sessions.
We digitized the calibration images and high speed videos, and used direct linear transformation
(DLT) to reconstruct three-dimensional kinematics from the two-dimensional kinematics from
both dorsal and lateral views. Digitization and DLT calculations were performed using custom
software (DLTcal5 and DLTdv5, Hedrick, 2008). Axes were set such that +x pointed in the
direction of forward motion, +z pointed vertically upward, and +y pointed to the left of the animal.
Footfall patterns of touchdown and takeoff were determined from the videos. On the granular
surface, because the hind foot often remained obscured by splashed grains during foot extraction,
we defined foot takeoff as when the knee began to flex following extension during limb
protraction (which is when foot takeoff occurs on the solid surface). To reduce noise and enable
direct comparisons among different running trials, position data were filtered with a Butterworth
low-pass filter with a cutoff frequency of 75 Hz, and interpolated to 0100% of one full stride
period (T) between two successive touchdowns of the right hind limb. All data analysis was
completed with MATLAB (MathWorks, Natick, MA, USA) unless otherwise specified.
Statistics
We accepted trials that met the following criteria: the animal ran continuously through the field of
view, the run was straight without contacting sidewalls of the trackway, there was a full stride
(between two consecutive touchdowns of the right hind limb) in the range of view, all the nine
markers were visible throughout the full stride, and the forward speed changed less than 20%
after the full stride. With these criteria, out of a total of 125 trials from 7 individuals on both solid
(61 trials from 7 individuals) and granular (64 trials from 7 individuals) surfaces collected over a
period of over three months, we ultimately accepted 51 runs from 7 individuals on solid (23 runs
from 7 individuals) and granular (28 runs from 7 individuals) surfaces. Because the data set had
an unequal number of runs per individuals, and because we were measuring freely-running
animals and did not control for speed, to maintain statistical power, all statistical tests were
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
6
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
performed on a subset of these data using one representative run per individual on both solid (N =
7) and granular (N = 7) surfaces. The representative run for each individual was selected based on
having the most consistent running speed for at least one full stride and was also closest to the
mean running speed of all 51 trials. Data are reported as mean ± 1 standard deviation (s.d.) from
the 7 representative runs on each substrate unless otherwise specified.
To determine the effect of substrate, all kinematic variables were corrected for size-related
differences by regressing the variables against SVL and taking the residuals for those that
regressed significantly with SVL (P < 0.05). We then ran an ANCOVA with substrate and speed
as covariates to test for substrate effects, independent of running speeds. All statistical tests were
performed using JMP (SAS, Cary, NC, USA).
For the energetics data, we used dimensionless quantities by normalizing energies of each run to
the CoM mechanical energy at touchdown of that run, thus eliminating the effect of mass and
running speed on energies. An ANOVA was used to test the differences between the reduction in
CoM mechanical energy, elastic energies, and energy loss. A Tukey’s HSD was used for post-hoc
tests where needed.
Dissection and model of hind limb
To gain insight into the role of anatomical components of the hind limb on mechanics during
locomotion, we dissected the hind limb of two preserved specimens. We quantified anatomical
dimensions by measuring the radii of the knee (K), ankle (A), the metatarsal-phalangeal joint
(MP), the distal end of the proximal phalanx (PP), and the digit tip (T) of the fourth toe. We also
observed the muscle and tendon arrangements within the lower leg and the foot. Based on these
anatomical features, we developed a model of the hind limb which incorporated the structure,
properties, and function of its main elements.
Resilience measurements of hind limb
To characterize the resilience of the hind limb for estimation of energy return, a modification of
the work loop technique was used (Fig. 2A), in which the limb was kept intact and forces were
applied to the whole limb instead of a single muscle (Dudek and Full, 2006). The animal was
anesthetized using 2% isoflurane air solution during the test. The hind foot was maintained within
the vertical plane, pushed down onto and then extracted from a custom force platform suited for
small animals (10.2 × 7.6 cm2, range = 2.5 N, resolution = 0.005 N) bonded with sandpaper (grit
size ~ 0.1 mm). Ground reaction force F was measured at 10 kHz sampling rate using a custom
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
7
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
LabVIEW program (National Instruments, Austin, TX, USA). A Phantom high speed camera
(Vision Research, Wayne, NJ, USA) simultaneously recorded deformation of the foot from the
side view at 250 frame/s (shutter time = 500 s). High-contrast markers (Wite-Out, Garden Grove,
CA, USA) were painted on the joints of the hind foot (A, MP, PP, T, and a point on the tibia
above the ankle). The ambient temperature was maintained at 35°C during the test.
Videos of foot deformation were digitized to obtain the angular displacement of the foot =
t, i.e., the change in the angle formed by the tibia and the foot (from the ankle to the digit tip
of the fourth toe) (Fig. 2A). Angular displacement was synchronized with the measured torque
about the ankle (calculated from the measured ground reaction force) to obtain a passive work
loop. The damping ratio of the hind limb, i.e., the percentage of energy lost within the hind limb
after loading and unloading, was calculated as the fraction of area within a work loop relative to
the area under the higher loading curve (Fung, 1993). Hind limb resilience, i.e., the percentage of
energy returned by the foot after loading and unloading, was one minus the damping ratio (Ker et
al., 1987; Dudek and Full, 2006). An ANOVA was used to test the effect of maximal torque,
maximal angular displacement, loading rate, and individual animal on hind limb resilience.
Granular penetration force measurements
While comprehensive force models are still lacking to calculate ground reaction forces during
locomotion granular media, a low speed penetration force model was previously used to explain
the locomotor performance of a legged robot on granular media (Li et al., 2009). Similarly, to
estimate the vertical ground reaction force on the lizard foot during running on the granular
surface, we measured the vertical force on a plate slowly penetrating vertically into the granular
substrate (Fig. 2B). Before each trial, a fluidized bed (area = 24 × 22 cm2) prepared the granular
substrate (depth = 12 cm) into a loosely packed state (volume fraction = 0.58) (for details, see
Maladen et al., 2009). A robotic arm (CRS robotics, Burlington, OT, Canada) pushed a
horizontally-oriented plate vertically downward at 0.01 m/s into the granular substrate to a depth
of 7.6 cm, and then extracted the plate vertically at 0.01 m/s. The force on the plate was measured
by a force transducer (ATI Industrial Automation, Apex, NC, USA) mounted between the robotic
arm and the plate at 100 Hz sampling rate using a custom LabVIEW program (National
Instruments, Austin, TX, USA). The depth of the plate was measured by tracking the position of
an LED light marker mounted on the robotic arm in side view videos taken by a Pike high speed
camera (Edmund Optics, Barrington, NJ, USA). Two thin aluminum plates of different area were
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
8
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
252
253
used (A1 = 7.6 × 2.5 cm2 and A2 = 3.8 × 2.5 cm2; thickness = 0.6 cm). Three trials were performed
for each plate.
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
Results
Performance and gait
representative runs on both substrates). Figure 4 shows average forward speed v(cid:3364) x,CoM, stride
On both solid and granular surfaces, the zebra-tailed lizard ran with a diagonal gait, a sprawled
limb posture, and lateral trunk bending (see Fig. 3 and Movies 1, 2 in supplementary material for
frequency f, and duty factor D of the entire data set (all symbols; 23 runs on the solid surface and
28 runs on the granular surface) and of the representative runs (filled symbols; N = 7 on the solid
(N = 7) surfaces. On both surfaces, v(cid:3364)x,CoM increased with f (Fig. 4A, P < 0.05, ANCOVA), and D
surface and N = 7 on the granular surface). Table 2 lists mean values and statistical results for all
decreased with v(cid:3364)x,CoM (Fig. 4B, P < 0.05, ANCOVA). D ≈ 0.45 on both surfaces resulting in an
the gait and kinematic variables from the representative runs for both solid (N = 7) and granular
v(cid:3364)x,CoM (P > 0.05, ANOVA) nor D (P > 0.05, ANCOVA) significantly differ between surfaces.
Average stride length = v(cid:3364)x,CoM/f was 15% shorter on the granular surface (P < 0.05, ANCOVA).
aerial phase of approximately 5% stride period T between alternating stances (Fig. 5A). Neither
Center of mass kinematics
The lizard displayed qualitatively similar center of mass oscillations during running on both
surfaces. The CoM forward speed vx,CoM (Fig. 5B) and vertical position zCoM (Fig. 5C) oscillated at
2f, dropping during the first half and rising during the second half of a stance, i.e., reaching
minimum at mid-stance and maximum during the aerial phase. The CoM also oscillated medio-
laterally at f (Fig. 5D). Throughout the entire stride, zCoM was significantly higher on the solid
surface (P < 0.05, ANCOVA). The CoM vertical oscillations zCoM and lateral oscillations yCoM
did not differ between substrates (P > 0.05, ANCOVA).
Hind foot, hind leg, and trunk kinematics
The lizard displayed distinctly different hind foot, hind leg, and trunk kinematics during running
on solid and granular surfaces (Figs. 3, 6). On the solid surface, the lizard used a digitigrade foot
posture (Fig. 3AE, solid line/curve). During the entire stride, the hind foot engaged the solid
9
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
surface only with the digit tips. At touchdown, the toes were straight and pointed slightly
downward. The touchdown foot angle (measured along the fourth toe) was touchdown = 12 ± 4°
relative to the surface (Fig. 3A,E; Fig. 6A, red). During stance, the long toes pivoted over the
stationary digit tips (Fig. 3AC, vertical dotted line shows zero displacement) and hyperextended
into a c-shape (Fig. 3B, solid curve). The foot straightened again at takeoff, pointing downward
and slightly backward (Fig. 3C, solid line), and then flexed during swing (Fig. 3D, solid curve).
On the granular surface, the lizard used a plantigrade foot posture (Fig. 3F,J, solid line). At
touchdown, the hind foot was nearly parallel with the surface, with the toes spread out and held
straight. In the vertical direction, the foot impacted the granular surface at speeds of up to 1 m/s.
The ankle joint slowed down to ~ 0.1 m/s within a few milliseconds following impact (a few
percent of stride period T) while the the foot started penetrating the surface. The touchdown foot
angle was touchdown = 4 ± 3° relative to the surface (Fig. 3J; Fig. 6A, blue), significantly smaller
than that on the solid surface (P < 0.05, ANCOVA). During stance, the entire foot moved
subsurface and was obscured (Fig. 3G). The ankle joint remained visible right above the surface
and moved forward by about a foot length (Fig. 3FH, vertical dotted line shows ankle
displacement). The foot was extracted from the substrate at takeoff, pointing downward and
slightly backward, and then flexed during swing (Fig. 3I, solid curve).
As a result of foot penetration on the granular surface, both the knee height zknee (Fig. 6B) and
pelvis height zpelvis (Fig. 6C) were lower on the granular surface (P < 0.05, ANCOVA). In addition,
on the granular surface, the knee moved downward by a larger vertical displacement zknee during
the first half of stance (P < 0.05, ANCOVA; Fig. 6B), while the knee joint extended by a larger
angle knee during the second half of stance (P < 0.05, ANCOVA; Fig. 6D). Throughout the
entire stride, the trunk was nearly horizontal on the solid surface (Fig. 3AD, dashed line), but
pitched head-up on the granular surface (Fig. 3FI, dashed line; Fig. 6E). On both surfaces, the
hind legs were sprawled at an angle of sprawl ≈ 40° during stance (Fig. 3; sprawl is defined as the
angle between the horizontal plane and the leg orientation in the posterior view). In most runs, the
tail was farther from the solid surface and closer to the granular surface (Fig. 3).
Hind limb anatomy
From morphological measurements (Table 1), the hind foot of the zebra-tailed lizard comprised
42% of the hind limb length, and the longest fourth toe alone accounted for 63% of the hind foot
length. These ratios are in similar range to previous observations (Irschick and Jayne, 1999a). The
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
10
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
slender foot had a cross-sectional radius of r = 0.501.25 mm tapering distally, with reducing
joint radii: rK = rA = 1.25 mm, rMP = 0.75 mm, rPP = rT = 0.50 mm.
Unlike many cursorial mammals whose ankle extensor muscles of the lower hind leg have long
tendons (Alexander, 2003), ankle extensor tendons are nearly non-existent in the zebra-tailed
lizard (Fig. 7A). Instead, layers of elongate tendons were found in both the dorsal and ventral
surfaces of the foot. Our anatomical description is focused on the ventral muscle and tendon
anatomy in the hind limb and terms given to muscles and tendons follow (Russell, 1993). A large,
tendinous sheath, the superficial femoral aponeurosis, originates from the femoro-tibial
gastrocnemius, stretches across the ventral surface of the foot, and inserts on the metatarsal-
phalangeal joints for digits III and IV. The superficial portion of the femoro-tibial gastrocnemius
muscle body extends to the base of the ankle, thereby rendering the human equivalent of the
ankle extensor tendons (i.e., the “Achilles” tendon) absent. Deep to the superficial femoral
aponeurosis lie the flexor digitorum brevis muscles (not shown) which control the flexion of each
of the digits. Tendons from the flexor digitorum longus muscle located on the lower hind leg run
deep to the flexor digitorum brevis muscle bodies, and extend to the tips of the digits. No
additional tendons are visible deep to the flexor digitorum longus tendons.
Hind limb model
Based on the observed muscle and tendon anatomy, we propose a two-dimensional strut-spring
model of the hind limb (Fig. 7B), which assumes isometric contraction for the lower leg muscles
and incorporates the spring nature of the foot tendons. This model is inspired from previous
observations in large running and hopping animals of the strut-like function of ankle extensor
muscles (Biewener, 1998a; Roberts et al., 1997) and spring-like function of ankle extensor
tendons (for a review, see Alexander, 2003). Rigid segments (Fig. 7B, dashed lines), which are
free to rotate about joints within a plane, represent the skeleton. The ankle extensor muscles in the
lower leg, which originate on the femur and run along the ventral side of the tibia, are modeled as
a rigid strut (muscle strut, Fig. 7B, blue line) that contracts isometrically during stance in running.
A linear spring (tendon spring, Fig. 7B, red line), which originates from the distal end of the
muscle strut and extends to the digit tip, models the elastic foot tendons. The muscle strut and
tendon spring are ventrally offset from the midline of the skeleton at each joint by respective joint
radii.
Hind limb resilience
11
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Representative passive work loops (Fig. 8AC) showed that torque was higher when the foot
was pushed down on the solid surface than when it was extracted, similar to previous
observations in humans (Ker et al., 1987) and cockroaches (Dudek and Full, 2006). Maximal
torque was positively correlated with maximal angular displacement (F1,62 = 64.3188, P < 0.001,
ANOVA). The kinks observed in the middle of the loading curve were due to the fifth toe
contacting the surface. Average hind limb resilience calculated from the work loops was R = 0.44
± 0.12 (Fig. 8DF, 3 individuals, 64 trials). R did not differ between individuals (F2,61 = 2.1025, P
= 0.1309, ANOVA), and did not depend on maximal torque (F1,62 = 0.5208, P = 0.4732, ANOVA;
Fig. 8D), maximal angular displacement F1,62 = 0.0164, P = 0.8987, ANOVA; Fig. 8E, or
average loading rate (F1,62 = 1.1228, P = 0.2934, ANOVA; Fig. 8F).
Hind foot curvature, tendon deformation, and tendon stiffness
The observed three-dimensional positions of the hind limb fit well to the two-dimensional hind
limb model (Fig. 9AD), and enabled calculation of the curvature, tendon deformation, and
tendon stiffness of the hind foot (see Appendix). Calculated hind foot curvature (Fig. 9E, solid
curve) showed that the hind foot hyperextended during stance (positive ) and flexed during
swing (negative ). The foot was straight at touchdown and shortly after takeoff ().
Calculated tendon spring deformation l(Fig. 9E, dashed curve) showed that the tendon spring
stretched during the first half and recoiled during the second half of stance. The estimated tendon
spring stiffness was k = 4.4 × 103 N/m (see Appendix).
Mechanical energetics on solid surface
Using the observed CoM and hind limb kinematics, calculated tendon spring stiffness and
deformation, and measured hind limb resilience, we examined the mechanical energetics of the
lizard running on the solid surface (Table 3, Fig. 9F). From the observed CoM kinematics, in the
first half of stance, the mechanical energy of the CoM (kinetic energy plus gravitational potential
energy) decreased significantly from Etouchdown = 1.00 ± 0.00 at touchdown to Emid-stance = 0.81 ±
0.08 at mid-stance (F2,18 = 12.2345, P = 0.0004, ANOVA, Tukey HSD). In the second half of
stance, the mechanical energy of the CoM recovered to Eaerial = 0.95 ± 0.10 at mid aerial phase,
not significantly different from Etouchdown (Tukey HSD). The reduction in CoM mechanical energy
in the first half of stance Emech = 0.19 ± 0.08 is the mechanical work needed per step on the solid
surface. Note that the energies of each run were normalized to Etouchdown of that run.
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
12
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
At mid-stance, the elastic energy stored in the tendon spring was Estorage = 0.18 ± 0.13 (calculated
2, see Appendix), not significantly different from Emech (F1,12 = 0.0475, P =
from 1/2 klmax
0.8312, ANOVA). Because hind limb resilience R = 0.44 ± 0.12, the elastic recoil of the foot
tendons returned an energy of Ereturn = REstorage = 0.08 ± 0.06, or 41 ± 33% of the mechanical work
needed per step (Emech) on the solid surface. We verified that foot flexion during swing induced
little energy storage (< 0.1 Estorage) because the hind foot was less stiff during flexion (0.7 × 103
N/m) than during hyperextension (4.4 × 103 N/m).
Granular penetration force model
Although little is known about the kinematics and mechanics of the complex limb intrusions
during legged locomotion on granular surfaces, we took inspiration from previous observations
that horizontal drag (Maladen et al., 2009) and vertical impact (Katsuragi and Durian, 2007)
forces in glass particles were insensitive to speed when intrusion speed was below approximately
0.5 m/s. Because the kinematics observed on the granular surface suggest that the vertical speeds
of most of the foot relative to the ground were below 0.5 m/s during most of the stance phase (see
Appendix), we assumed that the ground reaction forces on the lizard’s feet were also insensitive
to speed. This allowed us to use the vertical penetration force measured at 0.01 m/s to model and
estimate the vertical ground reaction forces on the lizard foot.
From the force data on both plates (Fig. 10), vertical ground reaction force Fz was proportional to
both penetration depth z and projected area A of the plate (area projected into the horizontal
plane). Fz was pointing upward during foot penetration, and pointing downward during foot
extraction and dropped by an order of magnitude. These measurements were in accord with
previous observations of forces on a sphere penetrating into granular media (Hill et al., 2005).
Furthermore, we estimated from free falling of particles under gravity that it would take longer
than the stance duration (45 ms) for the grains surrounding a penetrating foot to refill a hole
created by the foot of maximal depth (zmax = 1.0 cm, see Appendix). Thus we assumed that the
vertical ground reaction forces were negligible during foot extraction.
Fz = (cid:3420) (cid:2009) (cid:1878)(cid:1827), for increasing (cid:1878),
0, for decreasing (cid:1878),
Therefore, we approximate the vertical penetration force as:
where is the vertical stress per unit depth, which is determined by the properties of the granular
material and increases with compaction (Li et al., 2009). Fitting Fz = zA to the force data
(1)
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
13
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
during penetration over regions where the plate was fully submerged and far from boundary (Fig.
10, dashed lines), we obtained=× 105 N/m3 for loosely packed 0.27 ± 0.04 mm diameter
glass particles.
Vertical ground reaction force on granular surface
During a stance on the granular surface, the CoM vertical speed vz,CoM (calculated from zCoM) was
approximately sinusoidal (Fig. 11A, dashed curve). This implies that the Fz on a lizard foot must
be approximately sinusoidal. In addition, the foot was nearly horizontal at touchdown, but pointed
downward and slightly backward during takeoff. In consideration of the functional form of the
penetration force (Eqn. 1), we hypothesized that during stance the foot rotated subsurface by /2
in the sagittal plane (Fig. 11C), increasing foot depth z but decreasing projected foot area A, thus
resulting in a sinusoidal Fz which reaches a maximum at mid-stance before the foot reaches
largest depth (see Appendix). A sinusoidal Fz is also possible for a fixed projected foot area if the
foot maintains contact on solidified grains. However, this is unlikely considering that during
stance the ankle moved forward at the surface level by a foot length.
Assuming that during stance the hind foot rotated by /2 in the sagittal plane at a constant angular
velocity, the vertical ground reaction force that each foot generated was Fz = 5mg/9 sin10t/9T
(see Appendix). The net vertical acceleration due to this Fz and the animal weight mg was az =
Fz/m – g (Fig. 11B; solid and dashed curves are az from both hind feet, shifted from each other by
T/2). The CoM vertical speed vz,CoM predicted from the total az on both hind feet (Fig. 11A,
dashed curve) agreed with experimental observations (Fig. 11A, solid curve). The slight under-
prediction of the oscillation magnitudes of vz,CoM was likely due to an over-estimation of duty
factor on the granular surface. This is because Fz may have dropped to zero even before takeoff if
the foot started moving upward before takeoff (Fig. 10).
Mechanical energetics on granular surface
Using the measured CoM kinematics, assumed foot rotation, and calculated vertical ground
reaction force, we examined the mechanical energetics of the lizard running on the granular
surface (Table 3, Fig. 11D). In the first half of stance, the mechanical energy of the CoM
decreased significantly from Etouchdown = 1.00 ± 0.00 at touchdown to Emid-stance = 0.86 ± 0.09 at
mid-stance (F2,18 = 6.6132, P = 0.007, ANOVA, Tukey HSD). In the second half of stance, the
mechanical energy of the CoM recovered to Eaerial = 0.99 ± 0.10 at mid aerial phase, not
significantly different from Etouchdown (Tukey HSD). The reduction in CoM mechanical energy in
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
14
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
434
435
436
437
438
the first half of stance Emech = 0.14 ± 0.09 is the mechanical work needed per step on the
granular surface. By integration of Fz over vertical displacement of the foot during stance (see
Appendix), the energy lost to the granular substrate per step was estimated as Esubstrate = 0.17 ±
0.05, not significantly different from Emech (F1,12 = 0.4659, P = 0.5078, ANOVA). Note that the
energies of each run were normalized to Etouchdown of that run.
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
Discussion
Conservation of spring-mass-like CoM dynamics on solid and granular surfaces
The observed kinematics and calculated mechanical energetics demonstrated that the zebra-tailed
lizard ran like a spring-mass system on both solid and granular surfaces. On both surfaces, the
CoM forward speed (Fig. 5B), vertical position (Fig. 5C), and lateral position (Fig. 5D) displayed
oscillation patterns that are in accord with predictions from the Spring-Loaded Inverted Pendulum
(SLIP) model (Blickhan, 1989) and the Lateral Leg Spring (LLS) model (Schmitt et al., 2002).
The small relative oscillations of the CoM forward speed (i.e., vx,CoM/vx,CoM << 1) was expected
because the Froude number was large for the lizard (see Appendix). The substantial sprawling of
the legs contributed to the medio-lateral oscillatory motion of the animal. Furthermore, on both
surfaces, the mechanical energy of the CoM oscillated within a step, reaching minimum at mid-
stance and maximum during the aerial phase (Fig. 9F, 11D), a defining feature of spring-mass
like running (Blickhan, 1989; Schmitt et al., 2002).
To our knowledge, ours is the first study to quantitatively demonstrate spring-mass-like CoM
motion in lizards running on granular surfaces. Spring-mass-like CoM motion was previously
observed in other lizards and geckos running on solid surfaces (Farley and Ko, 1997; Chen et al.,
2006), but it was not clear whether energy-saving by elastic elements played an important role.
Hind foot function on solid surface: energy-saving spring
Our study is also the first to quantify elastic energy savings in foot tendons in lizards during
running on solid surfaces. The significant energy savings (about 40% of the mechanical work
needed per step) in the zebra-tailed lizard’s hind foot tendons is in a similar range to the energy
savings by ankle extensor tendons and digital flexor tendons and ligaments in larger animals
(Alexander, 2003), such as kangaroos (50%, Alexander and Vernon, 1975), wallabies (45%,
15
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Biewener et al., 1998), horses (40%, Biewener, 1998b), and humans (35%, with an additional 17%
from ligaments in the foot arch, Ker et al., 1987).
This is surprising considering that the elastic energy saving mechanism was previously thought
less important in small animals (e.g., 14% in hopping kangaroo rat of ~ 100 g mass, Biewener et
al., 1981). Because the tendons of small animals are “overbuilt” to withstand large stresses during
escape, during steady-speed locomotion these tendons usually experience stresses too small to
induce significant elastic energy storage and return (Biewener and Blickhan, 1988; McGowan et
al., 2008). We verified that for zebra-tailed lizards running at ~ 1 m/s the maximal stress in the
foot tendons is 4.3 MPa (see Appendix), well below the 100 MPa breaking stress for most
tendons (Kirkendall and Garrett, 1997).
The zebra-tailed lizard’s elongate hind foot and digitigrade foot posture on the solid surface may
be an adaptation for elastic energy saving during rapid locomotion. Like other iguanids (Russell,
1993), this lizard does not have substantial ankle extensor tendons as large animals do.
Nevertheless, elongation of foot tendons and a digitigrade posture enhance the hind foot’s energy
saving capacity by decreasing tendon stiffness and mechanical advantage (Biewener et al., 2004)
(see Appendix). A recent study also found significant energy savings (53%) by elongate foot
tendons in running ostriches (Rubenson et al., 2011). More generally, elongation of distal limb
segments such as legs, feet, and toes which possess tendons may be an adaptation for energy
saving during rapid locomotion. Indeed, many cursorial animals including mammals (Garland Jr.
and Janis, 1993), lizards (Bauwens et al., 1995), and dinosaurs (Coombs Jr., 1978) display
elongation of distal limb segments. Short fascicles and long tendons and ligaments are often
found in the ankle extensor muscles and digital flexor muscles in large cursorial ungulates such as
horses, camels, and antelopes (Alexander, 2003).
Solid surface model assumptions
Our estimates of elastic energy storage and return on the solid surface assume isometric
contraction of lower leg muscles. However, muscles have a finite stiffness and do lengthen by a
small amount under limb tension (Biewener, 1998a; Roberts et al., 1997). Despite this difference,
our estimates still hold, because in the latter case both lower leg muscles and foot tendons behave
like springs, and the total stiffness remains the same (since external force and total deformation
remain the same). In the case where the muscles actively shorten during stance and further
lengthen the tendons (which does positive mechanical work on the tendons), the energy storage
and return in the tendons would increase. However, the overall energy efficiency would decrease
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
16
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
(with everything else being the same), because apart from energy lost in tendon recoil, energy is
further lost in the muscles that perform the mechanical work, i.e., muscle work is more expensive
than tendon work (Biewener and Roberts, 2000).
In addition, the hind limb resilience obtained from anesthetized lizards was assumed to be a good
estimate for hind limb resilience in running lizards. This is based on our observations that
resilience was independent of torque, angular displacement, and loading rate, as well as previous
findings that the damping properties of animal limbs are largely intrinsic to their structure and
material properties (Weiss et al., 1988; Fung, 1993; Dudek and Full, 2006). Future studies using
techniques such as tendon buckles (Biewener et al., 1998), sonomicrometry (Biewener et al.,
1998), ultrasonography (Maganaris and Paul, 1999), and oxygen consumption measurement
(Alexander, 2003) during locomotion are needed to confirm this assumption.
Hind foot function on granular surface: dissipative, force-generating paddle
The similarity between the observed and predicted vz,CoM on the granular surface supports the
hypothesis of subsurface foot rotation. We speculate that on the granular surface the foot
functions as a “paddle” through fluidized grains to generate force. This differs from previous
observations of the utilization of solidification forces of the granular media in a legged robot (Li
et al., 2009; Li et al., 2010b) and sea turtle hatchlings (Mazouchova et al., 2010) moving on
granular surfaces.
As the zebra-tailed lizard’s hind foot paddles through fluidized grains to generate force, energy is
lost to the substrate because grain contact forces in granular media are dissipative (Nedderman,
1992). A large foot can reduce energy loss to the granular substrate compared to a small one,
much like large snowshoes used by humans can reduce energy cost for walking on snow (Knapik
et al., 1997). From our model of foot-ground interaction on the granular surface, for a given
animal (constant weight), energy loss to the substrate is proportional to foot penetration depth,
and thus inversely proportional to foot area and substrate strength (see Appendix).
Granular surface model assumptions
In our modeling of the foot-ground interaction on the granular surface using the penetration force
model, we made two assumptions. First, we assumed that the ground reaction forces were
insensitive to speed. This is true in the low speed regime (<0.5 m/s for our glass particles,
Maladen et al., 2009) where particle inertia is negligible and forces are dominated by particle
friction. Because friction is proportional to pressure, and pressure is proportional to depth (Hill et
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
17
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
al., 2005), granular forces in the low-speed regime are proportional to depth (Fz=zA),
analogous to the hydrostatic forces in fluids (Fz=gzA, i.e. buoyant forces due to hydrostatic
pressure).
Second, we used the vertical stress per unit depth determined from vertical penetration of a
horizontally oriented disc to estimate forces on the foot as it rotates subsurface. In this calculation,
the effective vertical stress per unit depthcosfoot (see Appendix) depended on foot orientation
via a simple relation cosfoot (because projected area A=Afootcosfoot; see Appendix), and not on
direction of motion. However, our recent physics experiments (Li et al., in preparation) suggest
that stresses in granular media in the low speed regime depend on both orientation and direction
of motion in a more complicated manner, and that cosfoot overestimates vertical stress per unit
depth for all foot orientations and directions of motion except when the foot is horizontal and
moving vertically downwards. Therefore, our model must be overestimating hydrostatic-like
forces, and there must be additional forces contributing to the lizard’s ground reaction forces.
We propose that these additional forces are likely from hydrodynamic-like inertial forces
resulting from the local acceleration of the substrate (particles) by the foot. Analogous to
hydrodynamic forces in fluids (Vogel, 1996), for an intruder moving rapidly in granular media,
the particles initially at rest in front of the intruder are accelerated by, and thus exert reaction
forces on, the intruder. Hydrodynamic-like forces at ~1 m/s can play an important role both in
impact forces on free falling intruders (Katsuragi and Durian, 2007; Goldman and Umbanhowar,
2008) and in legged locomotion of small lightweight robots (Qian et al., 2012). We note that the
foot rotation hypothesis should hold regardless, because hydrodynamic-like forces are also
proportional to projected area (Katsuragi and Durian, 2007).
However, we know too little about the lizard’s subsurface foot kinematics and the force laws in
the high-speed regime on an intruder being pushed in a complex path within granular media (not
simply a free-falling
intruder)
to more accurately calculate both hydrostatic-like and
hydrodynamic-like forces. Future x-ray high-speed imaging experiments (Maladen et al 2009;
Sharpe et al., 2012) will reveal how the lizard foot was moving subsurface. Further studies of
intrusion forces in granular media in both low-speed (Li et al., 2013) and high-speed regimes can
provide a more comprehensive understanding of ground reaction forces during legged locomotion
on granular surfaces and provide better estimates of foot penetration depth and energy loss.
Comparison to water-running in basilisk lizard
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
18
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
The rapid impact of the foot on the surface at touchdown and hypothesized subsurface foot
rotation appear kinematically similar to the slap and stroke phases of the basilisk lizards running
on the surface of water (Glasheen and McMahon, 1996a; Hsieh, 2003). For the zebra-tailed lizard
running on sand, both granular hydrostatic-like and hydrodynamic-like forces can contribute to
vertical ground reaction force. This is also qualitatively similar to water-running basilisk lizard,
which utilizes both hydrostatic forces resulting from the hydrostatic pressure between the water
surface and the bottom of the air cavity created by the foot, and hydrodynamic forces resulting
from the water being accelerated from rest by the rapidly moving foot (Glasheen and McMahon,
1996a, 1996b; Hsieh and Lauder, 2004).
However, the degree to which each species relies on these two categories of forces differs due to
differences in the properties of the supporting media. For given foot size, depth, and speed, the
hydrostatic(-like) forces in water are an order of magnitude smaller than the hydrostatic-like
forces in granular media, whereas the hydrodynamic(-like) forces are similar between in water
and in granular media (see Appendix). As a result, the basilisk lizard running on water must rely
on hydrodynamic forces to a larger degree than the zebra-tailed lizard running on sand,
considering that these two lizards have similar size (~ 0.1 m). An extreme example for this is that
it is impossible for a basilisk lizard to stand on the surface of water, but a zebra-tailed lizard can
stand on loose sand.
Motor function of upper hind leg
Despite the passive nature of the leg spring in the spring-mass model, animal limbs do not
function purely passively as springs—the muscles within them must perform mechanical work.
We have shown that on the solid surface, the lizard’s hind foot saves about 40% of the
mechanical work per step. The remaining 60% is lost either within the foot or to the ground, and
must be compensated by mechanical work performed by muscles, which is Wmuscle = 0.11 ± 0.10.
This work is likely provided by knee extension during the second half of stance (Fig. 6D, red
curve) powered by the upper leg muscles.
On the granular surface, substantial energy is lost to the substrate. This is in accord with previous
observations of higher mechanical energetic cost during locomotion on granular surfaces in
human (Zamparo et al., 1992; Lejeune et al., 1998) and legged robots (Li et al., 2010a). Because
the energy lost to the substrate equals the reduction in CoM mechanical energy during the first
half of stance, even without energy loss within the limb, the upper hind leg muscles must perform
mechanical work of Wmuscle = 0.31 ± 0.10 during the second half of stance, about three times that
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
19
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
on the solid surface for a given animal running at a given speed, as evidenced by the larger knee
extension on the granular surface (Fig. 6D, blue curve).
Our models of the foot-ground interaction on both surfaces assume purely passive foot mechanics,
and do not consider the role of active neurosensory control. However, animals can actively adjust
kinematics and muscle function to accommodate changes in surface conditions (Ferris et al., 1999;
Daley and Biewener, 2006). We observed that when confronted by a substrate which transitioned
from solid into granular (or vice versa), the lizard displayed partial adjustment of foot posture
during the first step on the new surface, followed by full adjustment during the second step.
Future studies using neuromechanics techniques, such as EMG (Biewener et al., 1998; Sponberg
and Full, 2008; Sharpe and Goldman, in review) and denervation and reinnervation (Chang et al.,
2009), can determine how neural control and sensory feedback mechanisms are used to control
limb function to accommodate changing substrates.
Conclusions
During running on both solid and granular surfaces, the zebra-tailed lizard displayed spring-mass-
like center of mass kinematics with distinct hind foot, hind leg, and trunk kinematics. The lizard’s
large, elongate hind foot served multiple functions during locomotion. On the solid surface, the
hind foot functioned as an energy-saving spring and reduced about 40% of the mechanical work
needed each footstep. On the granular surface, the hind foot paddled through fluidized grains to
generate force, and substantial energy was lost during irreversible deformation of the granular
substrate. The energy lost within the foot and to the substrate must be compensated for by
mechanical work done by the upper hind leg muscles.
The multifunctional hind foot may passively (and possibly actively) adjust to the substrate during
locomotion in natural terrain, and provide this desert generalist with energetic advantages and
simplify its neurosensory control tasks (Full and Koditschek, 1999). Current robotic devices often
suffer performance loss and high cost of transport on flowing substrates like granular material
(Kumagai 2004; Li et al., 2009; Li et al., 2010b; Li et al., 2010a). Insights from studies like ours
can provide inspiration for next-generation multi-terrain robots (Pfeifer et al., 2007). Finally, our
study also highlights the need for comprehensive force models for granular media (Li et al., in
preparation) and for flowing terrestrial environments in general.
Appendix
20
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Small relative oscillation in forward speed
2/gL0 = 3
Running at 1.1 m/s, the lizard’s Froude number in the sagittal plane was Fr = vx,CoM
(where L0 ≈ 4 cm is the leg length at touchdown), above the typical value of 2.5 where most
animals transition from trotting to galloping (Alexander, 2003). This implied that the kinetic
2 ≈ ½ mvx,CoM
2) of the CoM was 3 times larger than its gravitational potential
energy (½ mvCoM
energy (mgzCoM). Because both the forward speed oscillation vx,CoM and vertical speed oscillation
vz,CoM were determined by the total ground reaction force and the attack angle of the leg spring
( = sin-1 (vx,CoMDT/2L0) = 0.9 rad), they must be of the same order of magnitude (Blickhan, 1989),
i.e., vx,CoM ~ vz,CoM. From the observed CoM kinematics, vz,CoM < (gL0)1/2. Therefore, vx,CoM ~
vz,CoM < (gL0)1/2 << vx,CoM, and vx,CoM/vx,CoM << 1.
Hind foot curvature on solid surface
Three-dimensional kinematics showed that the hind limb (from the hip to the digit tip of the
fourth toe) remained nearly within a plane during the entire stride (out-of-plane component is 5%
averaged over the entire stride). During stance, the orientation of the foot plane remained nearly
unchanged, with a foot sprawl angle of 53 ± 4° relative to the sagittal plane in the posterior view.
Hind foot curvature could then be obtained by fitting a circle to the hind foot (from the ankle to
the digit tip) within the foot plane and determining the radius of curvature of the fit circle (see
diagram in Fig. 9A), i.e., = ± 1/, where + sign indicates foot hyperextension, sign indicates
foot flexion, and = 0 indicates a straight foot.
Tendon spring deformation
From the two-dimensional strut-spring model of the hind limb, by geometry, the tendon spring
deformation l was related to the observed changes of joint angles and the foot joint radii as: l =
irii, where i = K, A, MP, PP were the four joints in the modeli the observed changes of
joint angles, and ri the joint radii (rK = rA = 1.25 mm, rMP = 0.75 mm, rPP = 0.50 mm). We
observed that the relaxed hind foot of a live animal was nearly straight (Fig. 1A), which was
similar to the foot shape at touchdown during running (Fig. 3A,E). Thus we defined the relaxed
length of the tendon spring as the length when the foot was straight, i.e., l = 0 at touchdown.
Calculated maximal tendon spring deformation lmax = 0.78 mm corresponded to a 3% strain. We
did not consider tendon spring deformation in the swing phase (dotted curve in Fig. 6F) because
the assumption of isometric contraction of lower leg muscles was only valid for the stance phase.
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
21
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Tendon spring stiffness
The stiffness of the tendon spring was defined as the maximal tension divided by the maximal
deformation of the tendon spring, i.e., k = Tmax/lmax. From the observed CoM kinematics, the
total ground reaction force at mid-stance was Fmax = 0.3 N within the coronal plane and pointed
from the digit tip to the hip. At mid-stance, because the foot was neither dorsiflexing nor
plantarflexing, torque was balanced at the ankle, i.e., TmaxrA = FmaxxAT, where xAT = 1.4 cm was
the horizontal distance between the ankle and the digit tip at mid-stance, and rA = 1.25 mm. Thus
Tmax = 3.4 N and k = 4.4 × 103 N/m. The maximal stress in the foot tendons during stance was
2 = 4.3 MPa.
max = Tmax/rPP
The torsional stiffness of the ankle observed in anesthetized lizards from the modified work loop
experiments (~ 1 × 103 Nm/rad) was an order of magnitude smaller than estimated from running
kinematics (12 × 103 Nm/rad). This is however not contradictory but expected because during
stance the lizard’s lower leg muscles must be activated, and the resulting higher tension from
muscle contraction increases limb stiffness (Weiss et al., 1988).
Foot elongation increases energy savings on solid surface
The stiffness of a piece of elastic material like a tendon is k = E0A0/l0, where E0 is the Young’s
modulus (i.e., E0 is nearly constant). Thus, the stiffness of the tendon spring scales as k ∝ A0/l0 ∝
modulus, A0 the cross sectional area, and l0 the rest length of the material. Most animal tendons
r0
are primarily made of collagen (Kirkendall and Garrett, 1997) and are of similar Young’s
2/k ∝ 1/k for a given Tmax), an elongate tendon can store (and return) more energy.
2/l0, i.e., an elongate tendon (smaller r0 and larger l0) is less stiff and stretches more easily than
a short, thick tendon. Because elastic energy storage decreases with tendon stiffness (Estorage = ½
2 = ½ Tmax
klmax
An elongate foot also reduces the moment arm of tendon tension (small rA) but increases the
moment arm of the ground reaction force (large xAT) about the ankle, therefore reducing the
mechanical advantage (Biewener et al., 2004), so it increases tension in the foot for a given
ground reaction force (because Tmax = FmaxxAT/rA) and amplifies tendon stretch for enhanced
energy storage and return.
Vertical ground reaction force on granular surface
We assumed that the hind foot was rotating at a constant angular velocity about the moving
ankle during stance, i.e., foot = t within 0 ≤ t ≤ DT and 0 ≤ foot ≤ /2, then DT = 10/9T
22
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
= 35 rad/s. From the measured vertical speed of the ankle and this assumed foot rotation, the
vertical speed of most (75%) of the foot was always below 0.5 m/s during most (75%) of stance.
Given foot rotation foot = t, the foot area projected in the horizontal plane decreased with time
as A = Afootcost, where Afoot = 1 cm2 is the hind foot area; the foot depth (measured at the center
of the foot) increased with time as z = zmaxsint. The vertical ground reaction force on the foot
was
sinusoidal: Fz = Fz,maxsin2t, which was
then
sinusoidal, where Fz,max =
(cid:1516) (cid:1832)(cid:3053) ,(cid:3040)(cid:3028)(cid:3051) (cid:1871)(cid:1861)(cid:1866)2(cid:2033)(cid:1872) (cid:1856)(cid:1872)
Afootzmaxsin/4cos/4 = ½Afootzmax. For steady state locomotion on a level surface, the Fz
(cid:3021)(cid:2868)
generated by one foot averaged over a cycle must equal half the body weight, i.e.,
= ½mg. Therefore, Fz,max = 5mg/9 and Fz = 5mg/9 sin10t/9T.
Energy loss to granular substrate
(cid:3031) (cid:3053)(cid:3031)(cid:3047) (cid:1856)(cid:1872) , where zmax = 1.0 cm
(cid:1856) (cid:1878) = (cid:1516) (cid:1832)(cid:3053)(cid:3021)(cid:2868)
(cid:1832)(cid:3053)
loss to the granular substrate was Esubstrate = (cid:1516)
(cid:3053)(cid:3288)(cid:3276)(cid:3299)
(cid:2868)
By integration of vertical ground reaction force over vertical displacement of the foot, the energy
from Fz,max = ½Afootzmax. The hypothesized foot rotation in the sagittal plane did not take into
account the sprawl of the foot during stance, which could induce additional energy loss by lateral
displacement of the granular substrate. However, a sprawled foot posture did not affect the
condition of vertical force balance and thus did not change our estimate of energy dissipation in
the sagittal plane. Therefore this estimate provides a lower bound.
Large foot area reduces energy loss on granular surface
(cid:2033) (cid:1855)(cid:1867)(cid:1871)(cid:2033)(cid:1872) (cid:1856)(cid:1872) ∝ zmax ∝ 1/(Afoot). This implies that the energy loss to the granular
= (cid:1878)(cid:3040)(cid:3028)(cid:3051) (cid:1516) (cid:1832)(cid:3053)(cid:3021)(cid:2868)
For a given animal (constant weight mg), Fz,max = ½Afootzmax = 5mg/9 is constant, thus Esubstrate
substrate increases with foot penetration depth. On a given granular surface (fixed ), a larger
foot (larger Afoot) sinks less than a smaller foot, and thus loses less energy to the substrate. For a
given foot size (fixed Afoot), a foot sinks less on a stronger granular substrate (larger ) than on a
weaker substrate, and thus loses less energy to the substrate.
Comparison of forces in granular media and in water
For water, hydrostatic force is Fz = gzA. Comparing this with Fz = zA for granular media, g
is the equivalent of . For water, g = 1.0 × 104 N/m3; for loosely packed glass particles3.5
× 105 N/m3. Therefore, the hydrostatic forces in water are an order of magnitude smaller than the
hydrostatic-like forces in granular media for given foot size and depth.
23
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
709
710
711
712
713
Hydrodynamic(-like) forces should be proportional to the density of the surrounding media
because they are due to the media being accelerated. For water, = 1.0 × 103 N/m3; for loosely
packed glass particles the effective density is 2.5 × 103 N/m3 × 0.58 volume fraction = 1.45 × 103
N/m3. Therefore, the hydrodynamic forces in water and hydrodynamic-like forces in granular
media are on the same order of magnitude for given foot size and foot speed.
714
715
716
717
718
719
720
721
722
723
724
725
726
727
Acknowledgements
We gratefully thank Sarah Sharpe, Yang Ding, Nick Gravish, Ryan Maladen, Paul Umbanhowar,
Kyle Mara, Young-Hui Chang, Andy Biewener, Tom Roberts, Craig McGowan, and two
anonymous reviewers for helpful discussions and/or comments on the manuscript; Loretta Lau for
help with kinematics data tracking; Sarah Sharpe for help with animal protocol and
anesthetization; Mateo Garcia, Nick Gravish, and Andrei Savu for help with force plate setup;
Ryan Maladen and The Sweeney Granite Mountains Desert Research Center for help with animal
collection; and the staff of The Physiological Research Laboratory animal facility of The Georgia
Institute of Technology for animal housing and care. This work was funded by The Burroughs
Wellcome Fund (D.I.G. and C.L.), The Army Research Laboratory Micro Autonomous Systems
and Technology Collaborative Technology Alliance (D.I.G. and C.L.), The Army Research
Office Biological Locomotion Principles and Rheological Interaction Physics (D.I.G. and C.L.)
and The University of Florida and Temple University start-up funds (S.T.H.).
728
729
730
731
732
733
734
735
References
Alexander, R. M. (2003). Principles of Animal Locomotion. Princeton: Princeton University
Press.
Alexander, R. M. and Vernon, A. (1975). The mechanics of hopping by kangaroos
(Macropodidae). Journal of Zoology 177, 265–303.
Autumn, K., Liang, Y. A., Hsieh, S. T., Zesch, W., Chan, W. P., Kenny, T. W., Fearing, R. S.
and Full, R. J. (2000). Adhesive force of a single gecko foot-hair. Nature 405, 681–685.
24
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Bauwens, D., Garland Jr., T., Castilla, A. M. and Van Damme, R. (1995). Evolution of sprint
speed in lacertid lizards: morphological, physiological and behavioral covariation. Evolution 49,
848–863.
Biewener, A. A. (1998a). Muscle function in vivo: A comparison of muscles used for elastic
energy savings versus muscles used to generate mechanical power. Integrative & Comparative
Biology 38, 703–717.
Biewener, A. A. (1998b). Muscle-tendon stresses and elastic energy storage during locomotion in
the horse. Comparative Biochemistry and Physiology Part B: Biochemistry & Molecular Biology
120, 73–87.
Biewener, A. A., Alexander, R. M. and Heglund, N. C. (1981). Elastic energy storage in the
hopping of kangaroo rats (Dipodomys spectabilis). Journal of Zoology 195, 369–383.
Biewener, A. A. and Blickhan, R. (1988). Kangaroo rat locomotion: design for elastic energy
storage or acceleration? The Journal of Experimental Biology 140, 243–255.
Biewener, A. A., Farley, C. T., Roberts, T. J. and Temaner, M. (2004). Muscle mechanical
advantage of human walking and running: implications for energy cost. Journal of Applied
Physiology 97, 2266–2274.
Biewener, A. A., Konieczynski, D. D. and Baudinette, R. V. (1998). In vivo muscle force-
length behavior during steady-speed hopping in tammar wallabies. The Journal of Experimental
Biology 201, 1681–1694.
Biewener, A. A. and Roberts, T. J. (2000). Muscle and tendon contributions to force, work, and
elastic energy savings: a comparative perspective. Exercise and Sport Sciences Reviews 28, 99–
107.
Blickhan, R. (1989). The spring-mass model for running and hopping. Journal of Biomechanics
22, 1217–1227.
Blickhan, R. and Full R. J. (1993). Similarity in multilegged locomotion: Bouncing like a
monopode. Journal of Comparative Physiology 173, 509–517.
Chang, Y.-H., Auyang, A. G., Scholz, J. P. and Nichols, T. R. (2009). Whole limb kinematics
are preferentially conserved over individual joint kinematics after peripheral nerve injury. The
Journal of Experimental Biology 212, 3511–3521.
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
25
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Chen, J. J., Peattie, A. M., Autumn, K. and Full, R. J. (2006). Differential leg function in a
sprawled-posture quadrupedal trotter. The Journal of Experimental Biology 209, 249–259.
Coombs Jr., W. P. (1978). Theoretical aspects of cursorial adaptations in dinosaurs. The
Quarterly Review of Biology 53, 393–418.
765
766
767
768
769
Crawford, C. S. (1981). Biology of Desert Invertebrates. New York: Springer.
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
Daley, M. A. and Biewener, A. A. (2006). Running over rough terrain reveals limb control for
intrinsic stability. Proceedings of the National Academy of Sciences of the United States of
America 103, 15681–15686.
Dickinson, M. H., Farley, C. T., Full, R. J., Koehl, M. A. R., Kram, R. and Lehman, S.
(2000). How animals move: An integrative view. Science 288, 100–106.
Dickinson, W. W. and Ward, J. D. (1994). Low depositional porosity in eolian sands and
sandstones, Namib Desert. Journal of Sedimentary Research 64, 226–232.
Ding, Y., Gravish, N. and Goldman, D. I. (2011a). Drag induced lift in granular media.
Physical Review Letters 106, 028001(1–4).
Dudek, D. M. and Full, R. J. (2006). Passive mechanical properties of legs from running insects.
The Journal of Experimental Biology 209, 1502–1515.
Farley, C. T. and Ko, T. C. (1997). Mechanics of locomotion in lizards. The Journal of
Experimental Biology 200, 2177–2188.
Ferris, D., Liang, K. and Farley, C. T. (1999). Runners adjust leg stiffness for their first step on
a new running surface. Journal of Biomechanics 32, 787–794.
Ferris, D. P., Louie, M. and Farley, C. T. (1998). Running in the real world: Adjusting leg
stiffness for different surfaces. Proceedings of the Royal Society B: Biological Sciences 265, 989–
994.
Full, R. J. and Koditschek, D. E. (1999). Templates and anchors: neuromechanical hypotheses
of legged locomotion on land. The Journal of Experimental Biology 202, 3325–3332.
Fung, Y. C. (1993). Biomechanics: Mechanical Properties of Living Tissues. New York:
Springer.
26
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Garland Jr., T. and Janis, C. (1993). Does metatarsal/femur ratio predict maximal running
speed in cursorial mammals? Journal of Zoology 229, 133–151.
Glasheen, J. W. and McMahon, T. A. (1996a). A hydrodynamic model of locomotion in the
Basilisk Lizard. Nature 380, 340–341.
Glasheen, J. W. and McMahon, T. A. (1996b). Vertical water entry of disks at low Froude
numbers. Physics of Fluids 8, 2078–2083.
Goldman, D. I. and Umbanhowar, P. B. (2008). Scaling and dynamics of sphere and disk
impact into granular media. Physical Review E 77, 021308(1–14).
Gravish, N., Umbanhowar, P. B. and Goldman, D. I. (2010). Force and flow transition in
plowed granular media. Physical Review Letters 105, 208301(1–4).
Hedrick, T. L. (2008). Software techniques for two- and three-dimensional kinematic
measurements of biological and biomimetic systems. Bioinspiration & Biomimetics 3, 034001(1–
6).
Hill, G., Yeung, S. and Koehler, S. A. (2005). Scaling vertical drag forces in granular media.
Europhysics Letters 72, 137–143.
Holmes, P., Full, R. J., Koditschek, D. and Guckenheimer, J. (2006). The dynamics of legged
locomotion: Models, analyses, and challenges. SIAM Review 48, 207–304.
Hsieh, S. T. (2003). Three-dimensional hindlimb kinematics of water running in the plumed
basilisk lizard (Basiliscus plumifrons). The Journal of Experimental Biology 206, 4363–4377.
Hsieh, S. T. and Lauder, G. V. (2004). Running on water: Three-dimensional force generation
by basilisk lizards. Proceedings of the National Academy of Sciences of the United States of
America 101, 16784–16788.
Irschick, D. J. and Jayne, B. C. (1999a). Comparative three-dimensional kinematics of the
hindlimb for high speed bipedal and quadrupedal locomotion of lizards. The Journal of
Experimental Biology 202, 1047–1065.
Irschick, D. J. and Jayne, B. C. (1999b). A field study of the effects of incline on the escape
locomotion of a bipedal lizard, Callisaurus draconoides. Physiological and Biochemical Zoology
72, 44–56.
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
27
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Jindrich, D. L. and Full, R. J. (1999). Many-legged maneuverability: dynamics of turning in
hexapods. The Journal of Experimental Biology 202, 1603–1623.
Karasov, W. H. and Anderson, R. A. (1998). Correlates of average daily metabolism of field-
active zebra-tailed lizards (Callisaurus draconoides). Physiological Zoology 71, 93–105.
Katsuragi, H. and Durian, D. J. (2007). Unified force law for granular impact cratering. Nature
Physics 3, 420–423.
Ker, R. F., Bennett, M. B., Bibby, S. R., Kester, R. C. and Alexander, R. M. (1987). The
spring in the arch of the human foot. Nature 325, 147–149.
Kirkendall, D. T. and Garrett, W. E. (1997). Function and biomechanics of tendons.
Scandinavian Journal of Medicine & Science in Sports 7, 62–66.
Knapik, J., Hickey, C., Ortega, S., Nagel, J. and De Pontbriand, R. (2002). Energy cost
during locomotion across snow: a comparison of four types of snowshoes with showshoe design
considerations. Work 18, 171–177.
Korff, W. L. and McHenry, M. J. (2011). Environmental differences in substrate mechanics do
not affect sprinting performance in sand lizards (Uma scoparia and Callisaurus draconoides).
The Journal of Experimental Biology 214, 122–130.
Kumagai, J. (2004). Sand trip—DARPA’s 320-kilometer robotic race across the Mojave Desert
yields no winner, but plenty of new ideas. IEEE Spectrum 41, 44–50.
Lejeune, T. M., Willems, P. A. and Heglund, N. C. (1998). Mechanics and energetics of human
locomotion on sand. The Journal of Experimental Biology 201, 2071–2080.
Li, C., Hoover, A. M., Birkmeyer, P., Umbanhowar, P. B., Fearing, R. S. and Goldman, D. I.
(2010a). Systematic study of the performance of small robots on controlled laboratory substrates.
Proceedings of SPIE 7679, 76790Z(1–13).
Li, C., Umbanhowar, P. B., Komsuoglu, H. and Goldman, D. I. (2010b). The effect of limb
kinematics on the speed of a legged robot on granular media. Experimental Mechanics 50, 1383–
1393.
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
28
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Li, C., Umbanhowar, P. B., Komsuoglu, H., Koditschek, D. E. and Goldman, D. I. (2009).
Sensitive dependence of the motion of a legged robot on granular media. Proceedings of the
National Academy of Sciences of the United States of America 106, 3029–3034.
Li, C., Zhang, T. and Goldman, D. I. (2013). A terradynamics of legged locomotion on
granular media. Science, 339, 1408–1411.
Maganaris, C. N. and Paul, J. P. (1999). In vivo human tendon mechanical properties. Journal
of Physiology 521, 307–313.
Maladen, R. D., Ding, Y., Li, C. and Goldman, D. I. (2009). Undulatory swimming in sand:
Subsurface locomotion of the sandfish lizard. Science 325, 314–318.
Maladen, R. D., Ding, Y., Umbanhowar, P. B., Kamor, A. and Goldman, D. I. (2011).
Mechanical models of sandfish locomotion reveal principles of high performance subsurface
sand-swimming. Journal of The Royal Society Interface 8, 1332-1345.
Matson J. (2010). Unfree Spirit: NASA’s Mars rover appears stuck for good. Scientific American,
302, 16.
Mazouchova, N., Gravish, N., Savu, A. and Goldman, D. I. (2010). Utilization of granular
solidification during terrestrial locomotion of hatchling sea turtles. Biology letters 6, 398–401.
McGowan, C. P., Skinner, J. and Biewener, A. A. (2008). Hind limb scaling of kangaroos and
wallabies (superfamily Macropodoidea): Implications for hopping performance, safety factor and
elastic savings. Journal of Anatomy 212, 153–163.
Moritz, C. T. and Farley, C. T. (2003). Human hopping on damped surfaces: Strategies for
adjusting leg mechanics. Proceedings of the Royal Society B: Biological Sciences 270, 1741–
1746.
Mosauer, W. (1932). Adaptive convergence in the sand reptiles of the Sahara and of California:
A study in structure and behavior. Copeia 1932, 72–78.
Nedderman, R. M. (1992). Statics and Kinematics of Granular Materials. Cambridge:
Cambridge University Press.
Pfeifer, R., Lungarella, M. and Iida, F. (2007). Self-organization, embodiment, and
biologically inspired robotics. Science 318, 1088–1093.
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
29
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Qian F., Zhang T., Li C., Masarati, P., Hoover A.M., Birkmeyer P., Pullin A., Fearing R.S.
and Goldman D.I. (2012). Walking and running on yielding and fluidizing ground. Robotics:
Science and Systems (in press).
Roberts, T. J., Marsh, R. L., Weyand, P. G. and Taylor, C. R. (1997). Muscular force in
running turkeys: The economy of minimizing work. Science 275, 1113–1115.
Rubenson, J., Lloyd, D. G., Heliams, D. B., Besier, T. F. and Fournier, P. A. (2011).
Adaptations for economical bipedal running: The effect of limb structure on three-dimensional
joint mechanics. Journal of The Royal Society Interface 8, 740–755.
Russell, A. P. (1993). The aponeuroses of the lacertilian ankle. Journal of Morphology 218, 65–
84.
Schmitt, J., Garcia, M., Razo, R. C., Holmes, P. and Full, R. J. (2002). Dynamics and stability
of legged locomotion in the horizontal plane: A test case using insects. Biological Cybernetics 86,
343–353.
Sharpe, S. S., Ding, Y., and Goldman, D. I. (2012). Environmental interaction influences
muscle activation strategy during sand-swimming in the sandfish lizard (Scincus scincus). The
Journal of Experimental Biology 216, 260–274.
Spagna, J. C., Goldman, D. I., Lin, P.-C., Koditschek, D. E. and Full, R. J. (2007).
Distributed mechanical feedback in arthropods and robots simplifies control of rapid running on
challenging terrain. Bioinspiration & Biomimetics 2, 9–18.
Spence, A. J., Revzen, S., Seipel, J., Mullens, C. and Full, R. J. (2010). Insects running on
elastic surfaces. The Journal of Experimental Biology 213, 1907–1920.
Sponberg, S. and Full, R. J. (2008). Neuromechanical response of musculo-skeletal structures in
cockroaches during rapid running on rough terrain. The Journal of Experimental Biology 211,
433–446.
Vitt, L. J. and Ohmart, R. D. (1977). Ecology and reproduction of lower Colorado River lizards:
I. Callisaurus draconoides (Iguanidae). Herpetologica 33, 214–222.
Vogel, S. (1996). Life in Moving Fluids: The Physical Biology of Flow. Princeton: Princeton
University Press.
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
30
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Weiss, P. L., Hunter, I. W. and Kearney, R. E. (1988). Human ankle joint stiffness over the
full range of muscle activation levels. Journal of Biomechanics 21, 539–544.
Zamparo, P., Perini, R., Orizio, C., Sacher, M. and Ferretti, G. (1992). The energy cost of
walking or running on sand. European Journal of Applied Physiology and Occupational
Physiology 65, 183–187.
902
903
904
905
906
907
908
31
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
909
910
911
912
913
914
Fig. 1. Model organism and three-dimensional kinematics experiments. (A) A zebra-tailed lizard
resting on sand in the wild (photo: Thomas C. Brennan). (B) Experimental setup for three-
dimensional kinematics capture, with definitions of pelvis height (zpelvis), knee height (zknee), trunk
pitch angle (pitch), and knee angle (knee). Colored dots in (A,B) are digitized points on the
midline of the trunk, hind leg, and elongate hind foot.
32
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
915
916
917
918
919
920
921
Fig. 2. Experiments to measure hind limb resilience and granular penetration force. (A)
Experimental setup for hind limb resilience measurements. Dashed foot tracing shows the relaxed,
straight foot right before touchdown. Solid foot tracing shows the hyperextended foot during
ground contact. F, ground reaction force; 0, angle between the ankle and the digit tip in the
relaxed, straight foot; t, angle between the ankle and the digit tip in the hyperextended foot; ,
torque about the ankle. (B) Experimental setup for granular penetration force measurements.
33
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
922
923
924
925
926
927
Fig. 3. Lateral views of representative runs on the solid (AD) and the granular (FI) surface (see
Movies 1, 2 in supplementary material). (E,J) Closer views of foot posture at touchdown showing
definition of touchdown foot angle touchdown. Solid lines and curves along the foot indicate hind
foot posture and shape. Note that the lateral camera was oriented at an angle to the x, y, z axes
such that forward (+x) direction appeared to point slightly downwards.
34
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
928
929
930
931
932
933
Fig. 4. Performance and gait on the solid (red) and the granular (blue) surfaces. (A) Average
forward speed vs. stride frequency. (B) Duty factor vs. average forward speed. Different symbols
represent different individuals. Filled symbols are from the seven representative runs for each of
the seven individuals tested on both substrates. Empty symbols are from runs that were not
included in the representative data set.
35
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
934
935
936
937
Fig. 5. Center of mass (CoM) kinematics (mean ± s.d.) vs. time during a stride on the solid (red)
and the granular (blue) surfaces. (A) Footstep pattern. (B) CoM forward speed. (C) CoM vertical
position. (D) CoM lateral position. See Fig. 1 for definitions of kinematic variables.
36
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
938
939
940
941
942
Fig. 6. Hind foot, hind leg, and trunk kinematics (mean ± s.d.) vs. time during a stride on the solid
(red) and the granular (blue) surfaces. (A) Touchdown foot angle. (B) Knee height. (C) Pelvis
height. (D) Knee angle. (E) Trunk pitch angle. See Fig. 1 and Fig. 3E,J for definitions of
kinematic variables.
37
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
943
944
945
946
947
948
Fig. 7. Anatomy and a strut-spring model of the hind limb. (A) Ventral anatomy of a dissected
hind limb. Lower hind leg muscles are marked in blue; foot tendons are marked in red. (B) A
two-dimensional model of the hind limb. The muscle strut models isometrically contracting lower
leg muscles; the tendon spring models foot tendons. The radii of colored circles correspond to
measured joint radii.
38
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
Fig. 8. Hind limb resilience. (AC) Representative passive work loops of the hind foot (measured
at the digit tip) from each of the three anesthetized lizards tested. Different curves are from
different trials. The area within a work loop is the energy lost within the foot. See Fig. 2A for
schematic of experimental setup. (DF) Hind limb resilience vs. maximal torque, maximal
angular displacement, and average loading rate. Different symbols are from different individuals.
Solid and dashed lines in (DF) denote mean ± s.d.
949
950
951
952
953
954
955
956
957
39
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
958
959
960
961
962
963
964
965
966
967
968
969
Fig. 9. Foot-ground interaction on the solid surface. (AD) The hind foot shape from the lateral
view of a representative run on the solid surface. (AD) correspond with (AD) in Fig. 3. The
hind foot shape in the dorsal view is similar because the sprawl angle of the foot plane is nearly
constant during stance. The diagram in (B) defines the radius of curvature of the foot (see
Appendix). (E) Foot curvature (solid) and tendon spring deformation (dashed) (mean ± s.d.) vs.
time during a stride on the solid surface. Tendon spring deformation is not meaningful during
swing (dotted) when the muscle strut assumption does not hold. (F) Mechanical energies of the
CoM and elastic energies of the foot (mean ± s.d.) on the solid surface. All energies are
normalized to the mechanical energy of the CoM at touchdown (Etouchdown) for each run. *
indicates that Emid-stance is significantly different from Etouchdown and Eaerial (P < 0.05, ANOVA,
Tukey HSD).
40
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
970
971
972
973
Fig. 10. Granular penetration force (mean ± s.d.) vs. depth on two plates of different areas: A1 =
7.6 × 2.5 cm2 and A2 = 3.8 × 2.5 cm2. See Fig. 2B for schematic of experimental setup. Dashed
lines are linear fits to the data over steady state during penetration using Eqn. (1).
974
41
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
975
976
977
978
979
980
981
982
983
984
985
986
Fig. 11. Foot-ground interaction on the granular surface. (A) CoM vertical speed (mean ± s.d.) vs.
time during a stride. Solid curve is from experiment. Dashed curve is calculated from the vertical
acceleration from the model. (B) Vertical acceleration vs. time during a stride calculated from the
total vertical ground reaction force Fz on both feet and the animal weight mg. Solid and dashed
curves are the Fz on the two alternating hind feet. (C) Hypothesized subsurface foot rotation in the
sagittal plane. (FI) correspond with (FI) in Fig. 3. foot, foot angle in the vertical plane. (D)
Mechanical energy of the CoM and the energy loss to the substrate (mean ± s.d.) during running
on the granular surface. All energies are normalized to the mechanical energy of the CoM at
touchdown (Etouchdown) for each run. * indicates that Emid-stance is significantly different from
Etouchdown and Eaerial (P < 0.05, ANOVA, Tukey HSD).
42
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
987
988
Table 1. Morphological measurements (mean ± s.d.) of the seven individuals tested in the 3-D
kinematics experiments.
SVL (cm)
Mass m (g)
Trunk length (cm)
Pelvic width (cm)
Hind limb length (cm)
Hind foot length (cm)
Femur length (cm)
Tibia length (cm)
Tarsals and metatarsals length (cm)
Fourth toe length (cm)
7.2 ± 0.6
11.0 ± 2.7
4.4 ± 0.4
1.4 ± 0.1
6.4 ± 0.1
2.7 ± 0.1
1.6 ± 0.2
2.1 ± 0.2
1.0 ± 0.1
1.7 ± 0.1
989
43
990
991
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
F
0.4784
9.9101
0.5032
8.9112
5.4690
Table 2. Gait and kinematic variables (mean ± s.d.) and statistics using an ANCOVA. P values
reported are for substrate effect.
†Average forward speed v(cid:3364)x,CoM (m/s)
Variable
Average CoM height (cid:1878) CoM (cm)
Stride frequency f (Hz)
Duty factor D
Stride length (m)
Magnitude of CoM vertical oscillations
zCoM (cm)
Lowest CoM height zCoM (cm)
Time of lowest CoM height (T)
Highest CoM height zCoM (cm)
Average pelvis height pelvis (cid:1878) pelvis (cm)
Time of highest CoM height (T)
Average trunk pitch angle (cid:3364) pitch (deg)
Magnitude of CoM lateral oscillations
yCoM (cm)
Touchdown knee height zknee (cm)
Lowest knee height zknee (cm)
Knee vertical displacement during stance
zknee (cm)
Touchdown knee angle knee (deg)
Lowest knee angle knee (deg)
Highest knee angle knee (deg)
Knee joint extension during stance knee
(deg)
‡Average leg sprawl angle during stance
sprawl (deg)
Touchdown foot angle touchdown (deg)
Granular
Solid
1.1 ± 0.3
1.2 ± 0.3
7.5 ± 1.6
8.1 ± 2.0
0.46 ± 0.05 0.45 ± 0.07
0.16 ± 0.02 0.14 ± 0.02
2.2 ± 0.5
3.2 ± 0.7
0.3 ± 0.2
0.4 ± 0.3
3.0 ± 0.7
2.0 ± 0.4
0.18 ± 0.04 0.19 ± 0.04
3.3 ± 0.7
2.4 ± 0.6
0.44 ± 0.04 0.48 ± 0.01
0.86 ± 0.19 0.94 ± 0.23
3.1 ± 0.7
1.9 ± 0.5
9 ± 2
1 ± 3
1.7 ± 0.6
2.7 ± 0.7
1.8 ± 0.5
0.7 ± 0.4
1.1 ± 0.4
0.9 ± 0.2
88 ± 25
90 ± 13
79 ± 10
79 ± 17
150 ± 8
116 ± 15
37 ± 13
71 ± 4
3.7031
7.7544
0.9696
3.6126
3.0642
0.2350
8.8912
19.5282
6.7157
15.4261
0.7128
1.2344
1.3175
17.568
18.0994
P
0.5023
0.0319
0.5480
0.0409
0.0203
0.4697
0.0115
0.6366
0.0447
0.0325
0.5263
0.0046
0.0002
0.0171
0.0006
0.3056
0.6713
0.7549
0.0001
0.0001
N/A
0.0032
40 ± 1
12 ± 4
38 ± 5
4 ± 3
N/A
7.6973
992
All significant differences (P < 0.05) are in bold. Degree of freedom is (2,11) for all variables.
993
† An ANOVA was used to test the effect of substrate on running speed.
994
995
996
997
‡ A direct comparison was not possible for sprawl between substrates because sprawl was measured
differently: on the solid surface, leg orientation was measured from the hip to the digit tip; on the
granular surface, leg orientation was measured from the hip to the ankle.
44
Li et al. (2012), The Journal of Experimental Biology, 215, 3293–3308. doi:10.1242/jeb.061937
998
999
Table 3. Normalized energetic variables (mean ± s.d.). All energies were normalized to Etouchdown
for each run and averaged over 7 representative runs on each substrate.
Variable
Mechanical energy at touchdown Etouchdown
Mechanical energy at mid-stance Emid-stance
Mechanical energy during aerial phase Eaerial
Mechanical energy reduction Emech
Elastic energy storage at mid-stance Estorage
Elastic energy return Ereturn
Energy loss to substrate Esubstrate
Muscle mechanical work Wmuscle
Solid
1.00 ± 0.00
0.81 ± 0.08*
0.95 ± 0.10
0.19 ± 0.08
0.18 ± 0.13
0.08 ± 0.06
N/A
0.11 ± 0.10
Granular
1.00 ± 0.00
0.86 ± 0.09*
0.99 ± 0.10
0.14 ± 0.09
N/A
N/A
0.17 ± 0.05
0.31 ± 0.10
1000
1001
* indicates significant difference (P < 0.05) in the mechanical energy of the CoM at mid-stance
from that at touchdown and during aerial phase.
45
|
1806.04273 | 1 | 1806 | 2018-06-12T00:07:38 | Angiogenic Factors produced by Hypoxic Cells are a leading driver of Anastomoses in Sprouting Angiogenesis---a computational study | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.TO"
] | Angiogenesis - the growth of new blood vessels from a pre-existing vasculature - is key in both physiological processes and on several pathological scenarios such as cancer progression or diabetic retinopathy. For the new vascular networks to be functional, it is required that the growing sprouts merge either with an existing functional mature vessel or with another growing sprout. This process is called anastomosis. We present a systematic 2D and 3D computational study of vessel growth in a tissue to address the capability of angiogenic factor gradients to drive anastomosis formation. We consider that these growth factors are produced only by tissue cells in hypoxia, i.e. until nearby vessels merge and become capable of carrying blood and irrigating their vicinity. We demonstrate that this increased production of angiogenic factors by hypoxic cells is able to promote vessel anastomoses events in both 2D and 3D. The simulations also verify that the morphology of these networks has an increased resilience toward variations in the endothelial cell's proliferation and chemotactic response. The distribution of tissue cell`s and the concentration of the growth factors they produce are the major factors in determining the final morphology of the network. | physics.bio-ph | physics | OPEN
Received: 16 May 2017
Accepted: 29 May 2018
Published: xx xx xxxx
Angiogenic Factors produced by
Hypoxic Cells are a leading driver
of Anastomoses in Sprouting
Angiogenesis–a computational
study
Maurício Moreira-Soares
1, Rita Coimbra2, Luís Rebelo3, João Carvalho1 & Rui D. M. Travasso1
Angiogenesis - the growth of new blood vessels from a pre-existing vasculature - is key in both
physiological processes and on several pathological scenarios such as cancer progression or diabetic
retinopathy. For the new vascular networks to be functional, it is required that the growing sprouts
merge either with an existing functional mature vessel or with another growing sprout. This process
is called anastomosis. We present a systematic 2D and 3D computational study of vessel growth in
a tissue to address the capability of angiogenic factor gradients to drive anastomosis formation. We
consider that these growth factors are produced only by tissue cells in hypoxia, i.e. until nearby vessels
merge and become capable of carrying blood and irrigating their vicinity. We demonstrate that this
increased production of angiogenic factors by hypoxic cells is able to promote vessel anastomoses
events in both 2D and 3D. The simulations also verify that the morphology of these networks has an
increased resilience toward variations in the endothelial cell's proliferation and chemotactic response.
The distribution of tissue cells and the concentration of the growth factors they produce are the major
factors in determining the final morphology of the network.
Sprouting angiogenesis–a process by which new blood vessels grow from existing ones–is an ubiquitous phe-
nomenon in health and disease of higher organisms1, playing a crucial role in organogenesis, wound healing2,
inflammation3, as well as on the onset and progression of over 50 different diseases4 such as cancer5–7, rheumatoid
arthritis8 and diabetes9. This process leads to the formation of neo-vessel networks that irrigate tumors and other
hypoxic tissues. In the circumstances in which these neo-vessel networks are functional, they present a ramified
hierarchical tree-like structure at large scale, as observed in various tissues (from the mesenteric arteries, to the
retina)10–14, while at microvascular scale the capillaries form a lattice-like network connecting the arterial and
venous vascular trees10,14, as seen in reconstructed mouse vascular networks in ref.15 and used in a tumor model in
ref.16. To achieve this complex structure, the sprouting process involves different steps that are carefully regulated.
When tissue cells are in hypoxia, the Hypoxia Inducible Factor 1 α (HIF-1 α) accumulates and is translocated
to the cells' nucleus, where it works as a transcription factor for several genes encoding for proteins that take part
in preparing the cell's response to hypoxia (see review17 and also the mathematical models applied to inflam-
mation and cancer in refs18,19). The Vascular Endothelial Growth Factor (VEGF) is one of these proteins and is
necessary and sufficient to trigger angiogenesis. The VEGF secreted by hypoxic cells diffuses in the tissue, forming
well defined spatial concentration gradients. VEGF is present in different isoforms with different binding capacity
to the extracellular matrix (ECM) components20,21 (the specific shape of the gradient depends also on this bind-
ing ability to ECM components)22. At a nearby vessel, VEGF promotes the loosening of the adhesion between
vessel cells and triggers the activation of an endothelial cell (EC) that acquires the endothelial tip cell phenotype,
as modeled in23. This cell leads a new sprout and migrates in the direction of increasing VEGF concentration,
1CFisUC, Department of Physics, University of Coimbra, Coimbra, Portugal. 2Institute for Research in Biomedicine
(iBiMED), Department of Medical Sciences, University of Aveiro, Aveiro, Portugal. 3Department of Physics, Faculty
of Sciences, University of Lisbon, Lisbon, Portugal. Correspondence and requests for materials should be addressed
to M.M.-S. (email: [email protected]) or R.D.M.T. (email: [email protected])
1
SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8www.nature.com/scientificreportsas observed in in vivo mice retinas in ref.24. The tip cell state promotes the activation of the endothelial stalk cell
phenotype in the cells that follow its lead, as shown in ref.25, a research paper on the mouse retina. Proliferation
of the sprout stalk cells is triggered by both VEGF and the tension exerted by the tip cell, as predicted in a phase
field model in ref.26. In the new growing sprout the cells behind the tip cell are able to take its place, acquire the tip
cell phenotype for some tens of minutes, and drive sprout elongation, as displayed in studies in vitro and in vivo
of tip-stalk cells dynamics in mouse retinal vessels27,28. This dynamic behavior ensures that there is always a cell at
the front of the sprout with the tip phenotype capable of exerting a contractile force on the matrix, degrading and
remodeling matrix fibres and opening a pathway for the sprout to grow.
As the sprouts grow, ECs are able to either alter their shape or to extend and align vacuole in order to grow
a lumen connected to the initial vessel that is capable of carrying blood29,30, as shown in angiogenesis in vitro
in ref.31, in vivo in ref.32 and used in modeling in ref.33. However, in order for the blood to flow, it is required a
pressure gradient along the vessel, which is only achieved after the growing sprout merges either with an exist-
ing functional mature vessel or with another growing sprout that is in turn connected to an existing functional
mature vessel. The process by which sprouts meet and merge is called anastomosis, studied in mouse models
in34,35 and also discussed in the review36.
Though a pivotal step in angiogenesis, the mechanisms that drive vessel anastomosis are not yet completely
understood. In particular, there are several unanswered questions regarding how a sprout is guided in the direc-
tion of other vessels in the tissue. Macrophages and microglial cells play an important role in this process in vivo,
since they often sit at sites where vessels merge, stabilizing the connection between endothelial cells from different
sprouts34,37, demonstrated by in vivo experiments in the mouse retina in38. However, anastomoses still occur with-
out the presence of macrophages or microglial cells, though less frequently36,38.
Mechanical processes play an important role in driving vessel anastomosis. In fact, when a tip cell is next to an
existing vessel, its filopodia are able to sense their microenvironment, and to establish contact with either other
tip cells' filopodia or with other endothelial cells24,38,39, as modeled in refs36,40. This contact is then strengthened
by the VE-cadherins present in the filopodia41, the tip cell advances in the direction of the existing vessel and it
attaches to it39. The slow process of merging the two lumina at this new junction then ensues.
When the sprouts are at somewhat larger separations, mechanical factors are still able to play an important
role in the formation of anastomoses. Take for example the situation of two tip cells meeting directly. Tip cells
exert sufficient force to deform the matrix locally, and other tip cells are able to follow these deformation gradi-
ents26,40,42. In this way, a tip cell is able to sense the presence of other tip cells in its vicinity and will move in their
direction. However, this mechanism still functions at rather short ranges, on the order of the cell size, as shown
by in vitro experiments in ref.43, and used in agent-based models in refs42,44 only in the neighborhood of a tip cell
the tissue is sufficiently deformed to guide nearby tip cells.
In some tissues other type of mechanical cues also drive anastomoses events at larger ranges. In the retina, for
example, vessels grow above a layer of astrocytes that provide guidance and support24. In this situation, filopodia
mediate the contact between the endothelial cells and the astrocytes network underneath, guiding the neo-vessels
to almost replicate its structure. In this way, the astrocytes network plays an important role in driving vessel anas-
tomoses in the retina.
However, in tissues where there is no underlying cell network for vessels to replicate, and where they explore
the space between cells, we question how the vessels can be guided toward other vessels at larger scales in
order to be close enough for mechanical communication to be relevant. In this work, we will demonstrate that
pro-angiogenic factors, such as VEGF, play the decisive role in the process of anastomosis and in shaping the final
network. There is strong experimental data that supports this hypothesis. In sprouting angiogenesis, these factors
have a pivotal role in guiding the sprouts. Namely, VEGF in matrices is able to promote migration and anastomo-
ses24,45–47. So much so that VEGF micro-patterns in 3D matrices are able to guide mouse endothelial cells in vivo
to form anastomoses and functional vessels along those patterns48. Moreover, VEGF micro-beads also drive the
formation of functional vessels in their vicinity in vivo49,50.
In a non-pathological setting, hypoxic cells in vivo will be producing VEGF until they are irrigated, and they
can only be irrigated after there are anastomoses in their surroundings. Therefore VEGF will keep being produced
until nearby vessels merge and form loops capable of carrying blood. We will demonstrate that this mechanism
can also account for an increased resilience of the network morphology towards endothelial cell proliferation and
chemotactic response.
To address the capability of VEGF gradients to drive anastomoses in a neo-vessel network, we present an
extensive and systematic 2D and 3D computational study of vessel formation in a tissue. The 2D model can be
applied to angiogenesis in quasi two-dimensional vascular networks, such as networks in the retina, chick chorio-
allantoic membrane (CAM) assays, or other in vivo membranes. On the other hand, the 3D model can be applied
to angiogenesis in 3D scaffolds and in vivo tissues, as the adipose tissue, and in tumor angiogenesis.
In the last two decades angiogenesis has been modelled using different approaches. Initially, researchers
implemented continuous descriptions of angiogenesis that had the density of vessels as output51–56. These mod-
els were not aimed at obtaining the vascular network or the blood flow rate, but to provide an indication of the
advancement of the irrigated region. Angiogenesis was also modelled using more detailed discrete agent-based
models that take into account the interaction between every single cell in the system57,58. Discrete models have
been coupled to tumor growth models, used to describe neo-vascularisation in the retina, and to study in detail
the interaction between endothelial cells and their micro-environment59–65. Nevertheless, the large number of
parameters and rules often difficult the full exploration of the parameter space in these models. Recently, hybrid
models that combine a continuous description of the vessel sprout with a cell-based approach for tip cell crea-
tion and movement were developed22,26,66. These models allow the study of the shape of large vessel sprout net-
works with a smaller number of parameters. Nevertheless the role of chemical factors in regulating anastomoses
2
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8Figure 1. The value of the order parameters in different domains. The order parameter ψ is +1 inside the
tissue cells (represented in blue) and −1 outside, while the order parameter φ is +1 inside the endothelial cells
(represented in red) and −1 in the rest of the domain. The stroma is characterized by the region where φ and
ψ are both negative. The model dynamics prevents the superposition between the positive domains of the two
order parameters.
formation in 2D and in 3D has not yet been studied. For reviews on the literature of angiogenesis modelling see
for example67–71.
In the next section we will introduce the computational model used. In the Results and Discussion section we
will present and compare the results of the 2D and 3D models regarding the capability of VEGF to drive anasto-
moses formation. It will be stressed the role of this mechanism in determining the morphology of the vascular
networks, rendering it less dependent on endothelial stalk cell proliferation rates or on the migration velocity of
tip cells. Finally we draw the conclusions of the work in the last section.
Methods
The angiogenesis mathematical model implemented is a natural extension of the model introduced in22, and fur-
ther explored in40,72–74. Here, we will review, in an abbreviated way, this mathematical model underlining the two
major improvements that are introduced: the description of the cells in hypoxia using a second order parameter,
and the estimation of tissue irrigation to couple it with hypoxia regulation.
We model the ECs and the ECM through a phase-field formalism22,75,76, defining an order parameter, φ(r, t)
that identifies the vessel network: the value for this order parameter is approximately equal to 1 inside the vessel
network, and equal to −1 outside. In the present work, we introduce a second order parameter to describe the
tissue cells ψ(r, t), which is equal to 1 inside these cells and −1 outside. In this way, the vessels, ECM and tissue
cells can be unequivocally identified by the local values of φ and ψ (see Fig. 1).
The equation for the evolution of the order parameter φ(r, t) is composed of two main terms: one describing
the dynamics of the interface, and another describing the proliferation of endothelial cells (function of the angi-
ogenic factor concentration, T):
φ
2
∂ = ∇
t
M
− − ∇ +
(cid:31)
(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)
φ
φ
φ
ε
[
3
2
2
2
(
1) ]
γ φ
(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)
(cid:28)
(cid:30)
2)(
1)(
−
+
ψ
ψ
2
−
Interface Dynamics
+
T
(
( )
)
φ
φ
α
Θ .
p
(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)
(cid:28)
(cid:29)(cid:29)(cid:29)(cid:29)(cid:29)
(cid:30)
(cid:31)
EC Proliferation
(1)
The first term is directly obtained from the system's phenomenological free-energy functional (see
Supplementary Information S1), which includes a surface tension term22 and an energy cost for the overlap of
vessels and tissue cells77. The γ in (1) is proportional to this energy cost and is set high enough in order to avoid
appreciable overlap in our simulations. In equation (1) M is the mobility coefficient, ε is the width of the capillary
wall (proportional to the surface tension), and Θ(φ) is the Heaviside step function. The proliferation rate is rep-
p , for T < Tp, with βp constant; T being the angiogenic factor concentration,
resented by the function α
T
and Tp the angiogenic factor concentration for which the proliferation reaches its maximum value. For T > Tp we
set αp(T) to the constant value α
p p. Due to the Heaviside function, the proliferation only occurs inside
T
the capillary.
Often hypoxic cells have a low motility, especially if they are part of an epithelial tissue. Therefore, we consider
in the simulation that tissue cells are static and do not change their shape. Hence, the order parameter ψ will not
be altered during simulation. Nevertheless, the simulation methods used permit to introduce motility and defor-
mation on the hypoxic cells, which will be explored in future work.
β=T
β=T
The angiogenic factor is modeled as a diffusive field T, that is consumed by the ECs with the rate αT:
(
)
)
(
p
p
(2)
The concentration of angiogenic factor is kept at a constant value Ts at the center of all active hypoxic cells.
∂ = ∇ −
t
During the simulation, the sources will be deactivated as the vessel network grows and as they are irrigated.
α φ
T ( )
φ
Θ .
T
T
2
D T
3
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8Figure 2. Schematic representation of the two different rules for hypoxia regulation. In the simulation with
Rule 1 a cell in hypoxia will deactivate its production of VEGF when the blood circulation reaches a site at a
distance from the cell shorter than the oxygen diffusion length. In the model with Rule 2, the VEGF production
will cease if any vessel reaches a point closer to the cell than the oxygen diffusion length, independently of
whether there is blood circulating in that vessel or not. The endothelial cells are represented in red, the tip cells
are the yellow circles and the angiogenic factor is represented in light blue. The cells in hypoxia are represented
as darker blue circles. The arrows indicate the vessels with blood flow.
The endothelial tip cells are represented by agents with radius Rc. A new agent is introduced whenever there is
a minimum angiogenic factor concentration Tc, a minimum angiogenic factor gradient Gm, and if the tip cell can-
didate is located at a distance larger than 4Rc from all other tip cells. In this way, we are able to retrieve the typical
"salt and pepper" pattern of tip/stalk cells arising from the Delta-Notch signaling23,27,39. Some recent works78 have
modeled the Delta-Notch signaling pathway in the context of angiogenesis also taking into account the Jagged1
ligand, which contributes to the stabilization of the ECs' phenotypes78,79, possibly influencing the tip/stalk cell pat-
tern. The implementation of this more complex Delta-Jagged-Notch mechanism is out of the scope of the present
work but it is expected to be introduced in future model developments.
All the activated tip cells migrate by chemotaxis, and their velocity is aligned with the local angiogenic factor
v
GM, with χ constant, GM being the gradient mod-
gradient. The tip cell velocity is given by
, for ∇ <T
GM the tip cell velocity reaches its maximum value
ulus for which the maximum velocity is attained. For ∇ ≥T
. A tip cell is not able to enter inside a tissue cell (independently if the cell is in hypoxia or not):
G
T/
χ ∇ ∇
M
when the distance between the tip cell and a tissue cell is lower than 2Rc, the radial component of the velocity is
set to zero and the tip cell moves in the azimuthal direction around the tissue cell.
The value of φ inside the tip cell is set to a constant value that depends on the tip cell velocity and on the local
χ= ∇T
T
ECs' proliferation rate22.
Blood is able to flow once there are anastomoses in the vascular network. In order to estimate the irrigated
regions in the tissue, we first identify the vessels where the blood is flowing. We start by extracting the medial lines
of the vascular structure, thus finding the network bifurcation sites and the length of every vessel. The extraction
method used for this purpose is detailed in80. Afterwards, we calculate the blood flow rate in every vessel (see
Supplementary Information S1). In the simulation there is a single input node (at the top of the simulation box)
and a single exit node (at the bottom of the simulation box). For more details regarding the mathematical model
and for the values of all parameters please refer to the Supplementary Information S1.
Cells in Hypoxia - Rules 1 and 2. With the aim of addressing the capability of VEGF to drive anastomosis
formation, we compare the results of the model obtained with two alternative rules concerning the tissue angi-
ogenic factor production. When the new vessels are capable of delivering oxygen to the tissue, the newly irri-
gated cells deactivate their hypoxia mechanism and cease the production of angiogenic factors. Therefore several
models in the literature10,14,60,81 use estimates of tissue irrigation in order to determine which cells are producing
angiogenic factors. To mimic this mechanism we consider a simplified rule to identify the VEGF producing cells:
Rule 1: A cell produces angiogenic factor in the model if it is located at a distance larger than the average oxy-
gen diffusion length (d = 25 μm)82 from all vessels with blood flow rate different from zero (Fig. 2 left).
This is the simplest rule guaranteeing that anastomoses events are required for hypoxic cell deactivation.
However, many other models focus on the ability of the angiogenic factor to drive vessel growth and not anasto-
moses22,40,83,84. These models implement a different rule for tissue cell deactivation:
Rule 2: A cell produces angiogenic factor in the model if it is located at a distance larger than the average oxy-
gen diffusion length (d = 25 μm) from all vessels (Fig. 2 right).
In short, with Rule 2 chemical guidance promotes sprouting but is not associated with tissue irrigation, while
in Rule 1 it is. Comparing the morphology of the networks formed under these two rules, we show how important
it is for tissue irrigation to regulate VEGF production to create functional vascular networks in 2D and in 3D.
The calculation of tissue irrigation is in general quite complex, being dependent on the oxygen pressure at
every vessel, on vessel diameter, flow rate (function of the non-constant viscosity due to variable hematocrit
and the Fahrëus-Lindqvist effect), hematocrit level (which is altered at non-symmetrical bifurcations), oxygen
diffusion constant in the tissue and oxygen uptake by tissue cells85–87. However, in the present work a detailed
4
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8Figure 3. Vessel growth in two dimensions. Evolution of a vascular system with an initial capillary at the
bottom and a random distribution of hypoxic cells (green circles), which drive the chemotactic response by
ECs. Four time points are represented. The first row of figures correspond to the vessel network grown under
the Rule 1 for hypoxic cell deactivation, while the second row of images correspond to the vessel network grown
under Rule 2. On the final column the irrigated vessels are identified in red. The number of anastomoses and the
number of vessels with circulating blood are much higher when Rule 1 is implemented.
calculation of tissue irrigation will not be required, since our aim is to demonstrate that the need for anastomoses
formation to stop VEGF production in tissues (as in Rule 1) will determine several characteristics of the observed
vascular morphology. Rule 1 keeps the model general, with all anastomoses providing the same amount of irri-
gation to the tissue. With a full calculation of tissue irrigation we would observe that the irrigation provided by
each anastomosis would depend on its location within the vessel network. This systematic study of vessel density
in growing vasculatures as a function of the geometry of the initial network and of the oxygen pressure is very
relevant, but it is outside the scope of the present article.
In this way, the final vasculatures obtained by running the simulation with Rule 1 will be directly compared to
the simulations where the angiogenic factor drives growth but not anastomoses (using Rule 2). We will then draw
conclusions regarding the capacity of VEGF gradients to drive vessel anastomoses.
Results and Discussion
Anastomoses in 2D. Similarly to the setting in22 we start with a random distribution of hypoxic cells in a
simulation box of size 375 × 375 μm2. These hypoxic cells produce the angiogenic factor which diffuses slowly in
the tissue. When the angiogenic factor concentration at the vessel initially located at one edge of the box is larger
than Tc, a new tip cell is introduced in the simulation. This tip cell moves in the direction of the local angiogenic
factor gradient forming a new sprout. Progressively, new tip cells are activated and a ramified network is produced
(see Fig. 3). As the tip cells migrate in the tissue they are able to contour the hypoxic cells, diverting their move-
ment direction as they approach these cells.
The morphology of the obtained networks is strikingly dependent on the rule for hypoxic cell deactivation.
With Rule 2, the hypoxic cells start deactivating once the vessels pass nearby, leading to an open network with
fewer vessels and anastomoses. The vessels are straighter and directed towards the hypoxic tissue. As a result, only
few vessels are able to carry circulating blood. On the other hand, with Rule 1, the number of branches is higher.
The vessels surround several cells in hypoxia, irrigating them. The resulting vessel network morphology is more
lattice-like instead of tree-like. Many more vessels carry blood flow. The observed vasculature also suggests that
with Rule 1 regions with more hypoxic cells have a higher density of vessels, while regions with less cells have a
lower vessel density.
To verify how prevalent these differences between vessel morphology are, we carried out a systematic quanti-
tative study of the morphology obtained from our simulations in 2D and 3D. To characterize the vascular trees we
measure their branch density, average vessel diameter, and density of anastomoses. These descriptors were evalu-
ated as a function of the two model parameters that were considered most relevant for the network evolution22,88:
the maximum velocity of tip cell migration in the tissue (equal to χGM in the model, though only the value of χ
is altered), and the highest stalk cell proliferation rate (equal to βpTp in the model, only the value of βp is altered).
Notice that the proliferation term in the model corresponds to the increase of vessel area (or volume, in 3D) per
unit time: it is a measure of the change of space occupied by the vessel per unit time22. This description is ideal to
be directly compared with observations of the network evolution through diverse imaging techniques. However,
since the vessels are hollow, the cell proliferation in vivo is expected to be systematically smaller than the cell pro-
liferation (space increase) term in the model. Nevertheless, since the neo-vessels resulting from sprouting have
all similar calibers (see for example66 for in vivo results) and since also in our simulation the neo-vessels have all
approximately the width, the extra proliferation in our simulation serve simply to fill-in the lumen. Therefore, we
expect that, even taking into account the vessels lumen, the progress of the network morphology as a function of
the cells' proliferation rate would be the same as the one observed in our simulations.
The results of this analysis are presented in Fig. 4. Each data point in these plots represents the average value
from three random distributions of hypoxic cells. We start from a random distribution of hypoxic cells to avoid
5
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8Figure 4. Characterisation of the vascular networks in two dimensions. Comparison of the branches density
(A,D), anastomoses density (B,E) and average vessels diameter (C,F), between the two rules implemented in
the model, as a function of the maximum proliferation rate (top row) and as a function of the maximum tip cell
velocity (bottom row). The networks were analysed after they had grown sufficiently to deactivate the VEGF
production in all tissue cells. The error bars denote the error in the average of 3 simulations (for each condition).
any spatial bias. For both simulations with Rules 1 and 2, as the proliferation rate is increased, the vascular net-
works become more complex, with a larger number of vessels and anastomoses. With Rule 2, for very low prolif-
eration rates, the number of vessels is very small and there are no anastomoses (see Fig. 4A,B). As the proliferation
rate increases, the number of vessel branches increase rapidly until it reaches a high stationary number of about
700 branches per mm2. This increase in vessel branch density is accompanied by an increase in anastomoses den-
sity. The average vessel diameter increases only slightly with the stalk cell proliferation rate (Fig. 4C).
A similar qualitative behaviour regarding the number of vessel branches and anastomoses is observed with
Rule 1. However the values of these two descriptors are considerably higher. Remarkably, the vasculature mor-
phology is much more stable. In the range of proliferation explored in the simulation, we obtain an increase of
approximately ten-fold in the number of branches and anastomoses when using Rule 2 for hypoxic cell deacti-
vation, whereas in the simulation with Rule 1, this increase was only about two-fold. For most of the range of
proliferation rates explored, the average vessel diameter did not vary when Rule 1 was implemented.
Similar results were observed for variable tip cell velocity, namely, Rule 1 leads to a clear increase in the num-
ber of branches and anastomoses, but at the same time, to more stable vascular networks (Fig. 4D–F). For both
rules for hypoxic cell deactivation, as the tip cell velocity increases, the number of branches increases until a
maximum is reached. After this maximum, the velocity of the tip cell is so high that the sprouts are initially not
thick enough to accommodate tip cell activation (the vessels start being thinner than two cell radii). The activa-
tion of new tip cells can then only happen after there is enough proliferation in these thin vessels. Because of the
increased difficulty in tip cell activation, the number of branches decreases for high tip cell velocity. The density
of anastomoses in the network follows, qualitatively, the non-monotonic behaviour of the branch density. For
Rule 1 the number of branches and anastomoses are significantly higher and their relative variations with tip cell
velocity are significantly smaller (for example, with Rule 1 the number of anastomoses varies only about 2-fold
within the range of tip cells velocities explored, while with Rule 2 it varies about 15-fold within the same range).
With respect to vessel diameter, we observe that the vessels become thinner with increasing tip cell velocity until
they reach a constant value for tip cell velocities larger than the velocity where the maximum branch density is
attained. Figure 4F suggests a similar change in vessel diameter with tip cell velocity with both Rules 1 and 2.
We conclude that, in two dimensions, the hypoxic cell VEGF production levels can force anastomoses to be
created in the system. This mechanism also renders the morphology of the vascular networks much more insensi-
tive to the proliferation rate and tip cell velocity. The hypoxic cells are the main factor in determining the structure
of the vasculature that will satisfy their needs.
These conclusions are robust to the inclusion in the model of more realistic relations of, for example, the
chemotactic response on the degree of hypoxia in the cells. To explore the effect of variable chemotactic response,
we introduce a simplified dependence of χ in T, inspired in the experimental evidence that the expression levels
of VEGF receptor 2 in ECs are lower in hypoxia89,90. Hence, we considered a lower chemotactic response in this
case. We repeated the analysis reported in Fig. 4 with the chemotactic response as a function of the hypoxia level
6
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8Figure 5. Hypoxic cell distribution determine vascular network. Evolution of a vascular system with an initial
capillary at the bottom and uniformly distributed hypoxic cells (green dots), sources of VEGF which drive the
chemotactic response by ECs (maximum proliferation 2.8 × 10−4 s−1 and maximum tip cell velocity 0.375 μm/
min). Four time points are represented. The bottom row shows the same network growth with the vessels with
blood flow in red.
Figure 6. Neo-vessel morphology after 24 hours using the two rules. In the right panel (Rule 1) we mark the
regions of high vessel density, on the top and bottom, where the hypoxic cells are placed, in contrast with the
lower vessel density in a middle region (without hypoxic cells). With Rule 2 (left panel) the vessel density is
much more uniform throughout the domain.
− T
), and we did not observe any
)
(we used the local VEGF concentration, T, as proxy for hypoxia, χ
significant change in the observed trends (see Supplementary Information S1).
χ=
(1
0
To address the capability of hypoxic cell distribution to determine vascular structure, we run the two dimen-
sional model with Rule 1 for a regular distribution of hypoxic cells (see Fig. 5). In this simulation we set eight par-
allel rows of ten hypoxic cells with a large space between rows 4 and 5. The evolution of the system clearly shows
how the neo-vascular network grows and branches in order to feed the first rows of hypoxic cells. The branching
then almost stops when the capillaries cross the region without VEGF sources, going back to a very branched
morphology when they reach the second domain with hypoxic cells.
A late instant of the growth of this network is again depicted in Fig. 6 right, where it is evident that the density
of new vessels is higher, with more branches and anastomoses in the hypoxic regions, on the top and on the bot-
tom of the figure. On the region without cells in hypoxia (middle area) the branch density is lower and the num-
ber of anastomoses is reduced (for example, the branch density in this middle region is 0.81 × 103 mm−2; compare
to a branch density 1.8 × 103 mm−2 on the top hypoxic region of the system). Strikingly, in the corresponding
network obtained from the simulation with Rule 2 (see Fig. 6 left), the hypoxic cells do not affect dramatically the
vessel density, which is approximately constant throughout the simulated domain (with Rule 2 the branch density
in the middle region is 0.97 × 103 mm−2, similar to a branch density 1.1 × 103 mm−2 on the top hypoxic region).
Remarkably, using solely the production of chemical factors, and without including mechanical components,
we are able to guide the vasculature to form anastomoses where they are required. Similarly, sources of VEGF
in vivo are able to promote the anastomosis and formation of new vessels in their vicinity49,50. This result clearly
agrees with what is expected in angiogenesis due to hypoxia: a higher density of new vessels in the region with
more hypoxic cells. The vessel network is tailored to feed the tissue.
7
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8Figure 7. Vessel growth in three dimensions. Evolution of a vascular system with an initial capillary at
the center and randomly distributed hypoxic cells (green spheres), sources of VEGF, which drive the ECs
chemotactic response. Four time points are represented. The first row of images corresponds to the results for
the model under the Rule 1, while the bottom row to Rule 2. We clearly observe a higher density of vessels in the
simulation with Rule 1.
We conclude that, in two dimensions, the distribution of hypoxic cells and the production rate of angiogenic
factors are the main regulators of angiogenesis, not the proliferation of endothelial cells or the tip cell veloc-
ity. This conclusion raises questions regarding the power to change the vessel network by the newly proposed
anti-angiogenic and vessel normalization therapies that target the metabolism of endothelial cells in a way to
control their proliferation or migration velocity91–94. With these therapies, and without targeting VEGF producing
cells, only when these parameters are dramatically changed, the network morphology is critically altered (two
caveats: first, with lower tip cell velocity the vasculature would grow at a slower rate, but the number of branches
and anastomoses, when comparing networks of the same size, would not be significantly altered; second: the
simulation is run with the assumption that hypoxic cells greatly decrease their VEGF production rate once they
are irrigated, which does not happen, for example, in several high grade malignant tumors)4. However, the change
of endothelial cell proliferation or migration rates may have an important effect in the quality of the vessel (e.g.
lumen size, leakage, maturation)93.
However, vessels in most cases grow in three dimensional networks. Therefore, we extend the simulations to
3D to verify to which extent the conclusions regarding VEGF driven anastomoses hold in the formation of 3D
vascular networks.
Anastomoses in 3D.
In this subsection we are looking to verify if, like in two dimensions, the production of
angiogenic factors by the hypoxic cells is able, by itself, to guide vessels to meet each other in three dimensions.
We are also interested in understanding if the characteristics of endothelial cells, such as their proliferation rate
and tip cell velocity, have an important role in shaping these networks in three dimensions; or if, like in two
dimensions, modifications in these properties lead only to moderate changes in the vascular network.
The neo-vasculature in this three dimensional model is simulated in a box of size 250 × 250 × 125 μm3, it starts
from a central capillary72, and the new sprouts grow along the gradient of angiogenic factor, produced by hypoxic
cells distributed randomly in the matrix (see Fig. 7). As in the two dimensional model, a vessel is able to contour
a hypoxic cell it approaches.
As in two dimensions, the vessel network grown with Rule 1 is clearly more dense than the vessel network
obtained from the simulation with Rule 2 (see Fig. 7). At the end of the simulation with Rule 1 all the hypoxic cells
were irrigated, a consequence of the large number of anastomoses present in the network.
A systematic study of the density of branches, anastomoses and vessel diameter of the simulated three dimen-
sional networks was also carried out (see Fig. 8). Each data point in these plots represents the average value from
three random distributions of hypoxic cells.
Strikingly, anastomoses are almost absent in the simulations with Rule 2, independently of the proliferation
rate and tip cell velocity. This is different from what was observed in 2D, where Rule 2 was able to lead to the for-
mation of a large number of anastomoses for certain values of the parameters. In three dimensions, Rule 1 leads
to an increase in the branch density and to an impressive ten-fold increase in the density of anastomoses in the
network. Hence the production of VEGF until the tissue is irrigated is able by itself to guide the growing sprouts
toward each other, in order to form new anastomoses.
We also observe differences with the two dimensional scenario regarding the dependence of the network
morphology on the proliferation rate. Similarly to the two dimensional simulation, for larger endothelial cell
proliferation rate, the number of branches increases. However, in 3D this increase is smoother and spans a wider
range of proliferation rates. The increase on the number of branches with the proliferation rate occurs both in the
simulations with Rules 1 and 2. Regarding the density of anastomoses, we observe a plateau for higher prolifer-
ation rates. However, neither the number of branches nor the number of anastomoses can be said to be almost
independent on the proliferation rate, as it occurred in 2D.
8
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8Figure 8. Characterisation of the vascular networks in three dimensions. Comparison of the branches density
(A,D), anastomoses density (B,E) and average vessels diameter (C,F), between the two rules implemented on
the model, as a function of the maximum proliferation rate (top row) and as a function of the maximum tip cell
velocity (bottom row). The networks were analysed after they had grown sufficiently to deactivate the VEGF
production in all tissue cells. The error bars denote the error in the average of 3 simulations (for each condition).
On the other hand, when varying tip cell velocity, the relative change in network morphology parameters is
much smaller for both Rules 1 and 2. Though for lower tip cell velocities the vessel network grows slower, and
thus takes a longer time to reach the stationary state at the end of the simulation their final morphology is rather
independent of the value of the maximum tip cell velocity value.
The diameter of the vessels is not altered considerably neither with the increase in proliferation nor in tip cell
velocity. With the increase in proliferation there are simply more vessels, rather than thicker vessels. In the case
of increasing tip cell velocity, the network is rather stable, with the number of branches, anastomoses and branch
diameter not varying appreciably.
Conclusion
In this article we presented a model of vessel sprouting that addresses the VEGF gradients' role in driving the
formation of anastomoses. This important biological mechanism is fundamental to irrigate tissue cells. However,
there are still several unanswered questions regarding how a sprout is guided in the direction of other vessels in
the tissue. The role of mechanical cues in anastomose formation has been explored in the literature24,26,36,38–40,42,43
but it can only account for guidance at short distances between merging sprouts. In the present work we demon-
strated how slowly diffusible growth factors gradients, maintained by tissue cells until they have access to suffi-
cient oxygen and nutrients, are able to guide vessel sprouts toward each other at larger scales. This mechanism is
particularly relevant in three dimensions, where without it the network has a reduced number of anastomoses,
and it is not capable of irrigating the tissue.
The role of hypoxic cells is therefore pivotal in determining the final morphology of the network. We showed
that it is the location of these cells and the concentration of the growth factors they produce that better determine
the number of vessels and of anastomoses in the final network. Actually, these factors are more relevant to the
final morphology than the tip cell chemotactic response or the ECs' proliferation rate. Experimentally it was
shown the extraordinary capacity for growth factors spatial distributions in vivo to influence sprouting48,95 and to
determine the site of functional neo-vessels49.
This resilience of the final network morphology toward EC's properties in 2D and in 3D (to a smaller degree)
guarantees reliable vascular structures in non-pathological settings. However, it raises questions regarding the
capacity for modifying the vessel network of the newly proposed vessel normalization therapies, which control
EC's metabolism91,92 in tissues where cells are still able to regulate their VEGF production. With these therapies,
and without targeting VEGF producing cells, the network morphology is only critically altered when the EC's
proliferation rate or migration velocity are dramatically affected. In fact, in agreement with our work, it has been
observed experimentally in vivo that a reduction in the EC's proliferation rate of over 50% does not significantly
alterthe density of tumor vessel networks93. Moreover, to observe a dramatic change of the vasculature, one has to
produce a substantial change in EC's metabolism96, leading to very low proliferation rate and migration velocity,
just as predicted by our work. However, a moderate decrease in EC's proliferation is able to affect the capacity
9
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8of the network to irrigate the tissue due to a lower vessel leakage and an increase in lumen size91–94. However, in
conditions where tissue cells produce VEGF at permanently high rates independently of the irrigation provided
by the network, both the ECs' proliferation rate and the tip cell velocity are able to shape the vascular structure88.
In this mathematical model, the hypoxic cells function as sites for anastomoses formation, producing angio-
genic factors until vessels merge. This suggests that in vivo cells that promote anastomosis, such as macrophages
and microglia34,37,38 can fulfil this role by chemical guidance (from VEGF and other chemical factors). It is also
evident the relevance of targeting macrophage metabolism in anti-angiogenic therapy97.
References
1. Folkman, J. Angiogenesis in cancer, vascular, rheumatoid and other disease. Nature medicine 1, 27–30 (1995).
2. Tonnesen, M. G., Feng, X. & Clark, R. A. Angiogenesis in wound healing. Journal of Investigative Dermatology Symposium
Proceedings, vol. 5, pp. 40–46, Elsevier (2000).
FASEB Journal 11, 457–465 (1997).
3. Jackson, J. R., Seed, M., Kircher, C., Willoughby, D. & Winkler, J. The codependence of angiogenesis and chronic inflammation. The
4. Carmeliet, P. Angiogenesis in health and disease. Nature medicine 9, 653 (2003).
5. Liao, D. & Johnson, R. S. Hypoxia: a key regulator of angiogenesis in cancer. Cancer and Metastasis Reviews 26, 281–290 (2007).
6. Fidler, I. J. Angiogenesis and cancer metastasis. Cancer journal (Sudbury, Mass.) 6, S134–41 (2000).
7. Carmeliet, P. & Jain, R. K. Angiogenesis in cancer and other diseases. Nature 407, 249–257 (2000).
8. Maruotti, N., Cantatore, F., Crivellato, E., Vacca, A. & Ribatti, D. Angiogenesis in rheumatoid arthritis. Histology and histopathology
9. Martin, A., Komada, M. R. & Sane, D. C. Abnormal angiogenesis in diabetes mellitus. Medicinal research reviews 23, 117–145 (2003).
10. Secomb, T. W., Alberding, J. P., Hsu, R., Dewhirst, M. W. & Pries, A. R. Angiogenesis: an adaptive dynamic biological patterning
11. Guibert, R., Fonta, C. & Plouraboué, F. A new approach to model confined suspensions flows in complex networks: application to
problem. PLoS Comput Biol 9, e1002983 (2013).
blood flow. Transport in porous media 83, 171–194 (2010).
12. Masters, B. R. Fractal analysis of the vascular tree in the human retina. Annu. Rev. Biomed. Eng. 6, 427–452 (2004).
13. Flores, J., Poiré, E. C., Del Ro, J. & de Haro, M. L. A plausible explanation for heart rates in mammals. Journal of theoretical biology
(2006)
14. Gandica, Y., Schwarz, T., Oliveira, O. & Travasso, R. D. Hypoxia in vascular networks: a complex system approach to unravel the
15. Blinder, P. et al. The cortical angiome: an interconnected vascular network with noncolumnar patterns of blood flow. Nature
265, 599–603 (2010).
diabetic paradox. PLoS ONE 9, e113165 (2014).
neuroscience 16, 889–897 (2013).
Journal of mathematical biology 58, 689 (2009).
1163–1177 (2003).
e1004436 (2015).
943–953 (2010).
138, 4763–4776 (2011).
222–231 (2009).
16. Owen, M. R., Alarcón, T., Maini, P. K. & Byrne, H. M. Angiogenesis and vascular remodelling in normal and cancerous tissues.
17. Semenza, G. L. Hypoxia-inducible factors in physiology and medicine. Cell 148, 399–408 (2012).
18. Cavadas, M. A., Nguyen, L. K. & Cheong, A. Hypoxia-inducible factor (hif) network: insights from mathematical models. Cell
Communication and Signaling 11, 42 (2013).
19. Nguyen, L. K. et al. A dynamic model of the hypoxia-inducible factor 1α (HIF-1α) network. J Cell Sci 126, 1454–1463 (2013).
20. Ferrara, N., Gerber, H.-P. & LeCouter, J. The biology of vegf and its receptors. Nature medicine 9, 669–676 (2003).
21. Chen, T. T. et al. Anchorage of vegf to the extracellular matrix conveys differential signaling responses to endothelial cells. The
Journal of Cell Biology 188, 595–609 (2010).
22. Travasso, R. D. M., Corvera Poiré, E., Castro, M., Rodríguez-Manzaneque, J. C. & Hernández-Machado, A. Tumor angiogenesis and
vascular patterning: A mathematical model. PLoS ONE 6, e19989 (2011).
23. Bentley, K., Gerhardt, H. & Bates, P. A. Agent-based simulation of notch-mediated tip cell selection in angiogenic sprout
initialisation. Journal of Theoretical Biology 250, 25–36 (2008).
24. Gerhardt, H. et al. Vegf guides angiogenic sprouting utilizing endothelial tip cell filopodia. The Journal of Cell Biology 161,
25. Hellstrom, M. et al. Dll4 signalling through notch1 regulates formation of tip cells during angiogenesis. Nature 445, 776–780 (2007).
26. Santos-Oliveira, P. et al. The force at the tip - modelling tension and proliferation in sprouting angiogenesis. PLoS Comput Biol 11,
27. Jakobsson, L. et al. Endothelial cells dynamically compete for the tip cell position during angiogenic sprouting. Nat Cell Biol 12,
28. Arima, S. et al. Angiogenic morphogenesis driven by dynamic and heterogeneous collective endothelial cell movement. Development
29. Lubarsky, B. & Krasnow, M. A. Tube morphogenesis: making and shaping biological tubes. Cell 112, 19–28 (2003).
30. Iruela-Arispe, M. L. & Davis, G. E. Cellular and molecular mechanisms of vascular lumen formation. Developmental Cell 16,
31. Meyer, G. T., Matthias, L. J., Noack, L., Vadas, M. A. & Gamble, J. R. Lumen formation during angiogenesis in vitro involves
phagocytic activity, formation and secretion of vacuoles, cell death, and capillary tube remodelling by different populations of
endothelial cells. The Anatomical Record 249, 327–340 (1997).
32. Gebala, V., Collins, R., Geudens, I., Phng, L.-K. and Gerhardt, H. Blood flow drives lumen formation by inverse membrane blebbing
33. Boas, S. E. & Merks, R. M. Synergy of cell–cell repulsion and vacuolation in a computational model of lumen formatio. n. Journal of
during angiogenesis in vivo. Nature cell biology (2016)
The Royal Society Interface 11, 20131049 (2014).
tip cell induction. Blood 116, 829–840 (2010).
90, 1–11 (2013).
34. Fantin, A. et al. Tissue macrophages act as cellular chaperones for vascular anastomosis downstream of vegf-mediated endothelial
35. Zhan, K., Bai, L. & Xu, J. Role of vascular endothelial progenitor cells in construction of new vascular loop. Microvascular Research
36. Geudens, I. & Gerhardt, H. Coordinating cell behaviour during blood vessel formation. Development 138, 4569–4583 (2011).
37. Baer, C., Squadrito, M. L., Iruela-Arispe, M. L. & De Palma, M. Reciprocal interactions between endothelial cells and macrophages
in angiogenic vascular niches. Experimental Cell Research 319, 1626–1634 (2013).
38. Rymo, S. F. et al. A two-way communication between microglial cells and angiogenic sprouts regulates angiogenesis in aortic ring
cultures. PLoS ONE 6, e15846– (2011).
angiogenesis. PLoS Comput Biol 5, e1000549 (2009).
Journal of The Royal Society Interface, 14 (2017)
39. Bentley, K., Mariggi, G., Gerhardt, H. & Bates, P. A. Tipping the balance: Robustness of tip cell selection, migration and fusion in
40. Vilanova, G., Colominas, I. and Gomez, H. A mathematical model of tumour angiogenesis: growth, regression and regrowth.
41. Almagro, S. et al. The motor protein myosin-x transports ve-cadherin along filopodia to allow the formation of early endothelial
cell-cell contacts. Molecular and Cellular Biology 30, 1703–1717 (2010).
10
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8 42. van Oers, R. F. M., Rens, E. G., LaValley, D. J., Reinhart-King, C. A. & Merks, R. M. H. Mechanical cell-matrix feedback explains
pairwise and collective endothelial cell behavior in vitro. PLoS Comput Biol 10, 1–14 (2014).
43. Reinhart-King, C. A., Dembo, M. & Hammer, D. A. Cell-cell mechanical communication through compliant substrates. Biophysical
Journal 95, 6044–6051 (2008).
44. Ramos, J. R., Travasso, R. & Carvalho, J. Capillary network formation from dispersed endothelial cells: Influence of cell traction, cell
adhesion, and extracellular matrix rigidity. Physical Review E 97, 012408 (2018).
45. Shamloo, A., Ma, N., Poo, M. M., Sohn, L. L. & Heilshorn, S. C. Endothelial cell polarization and chemotaxis in a microfluidic device.
Lab on a chip 8, 1292–1299 (2008).
46. Benitez, P. & Heilshorn, S. Microfluidic devices for quantifying the role of soluble gradients in early angiogenesis. In Mechanical and
Chemical Signaling in Angiogenesis (pp. 47–70). Springer Berlin Heidelberg. (2013)
47. Koch, S. et al. Enhancing angiogenesis in collagen matrices by covalent incorporation of VEGF. Journal of Materials Science:
Materials in Medicine 17, 735–741 (2006).
48. Oh, H. H., Lu, H., Kawazoe, N. & Chen, G. Spatially guided angiogenesis by three-dimensional collagen scaffolds micropatterned
with vascular endothelial growth factor. Journal of Biomaterials Science 23, 2185–2195 (2012).
49. Elçin, Y. M., Dixit, V. & Gitnick, G. Extensive in vivo angiogenesis following controlled release of human vascular endothelial cell
growth factor: implications for tissue engineering and wound healing. Artificial Organs 25, 558–565 (2001).
50. Tufró, A. VEGF spatially directs angiogenesis during metanephric development in vitro. Developmental Biology 227, 558–566
51. Orme, M. E. & Chaplain, M. A. J. A mathematical model of vascular tumour growth and invasion. Math Comp Model 23, 43–60
52. Orme, M. E. & Chaplain, M. A. J. Two-dimensional models of tumour angiogenesis and anti-angiogenesis strategies. Math Med Biol
(2000).
(1996).
14, 189–205 (1997).
53. Levine, H. A., Sleeman, B. D. & Nilsen-Hamilton, M. Mathematical modeling of the onset of capillary formation initiating
angiogenesis. J Math Biol 42, 195–238 (2001).
54. Levine, H. A., Pamuk, S., Sleeman, B. D. & Nilsen-Hamilton, M. Mathematical modeling of capillary formation and development in
tumor angiogenesis: penetration into the stroma. Bull Math Biol 63, 801–863 (2001).
55. Levine, H. A., Sleeman, B. D. & Nilsen-Hamilton, M. A mathematical model for the roles of pericytes and macrophages in the
initiation of angiogenesis. I. The role of protease inhibitors in preventing angiogenesis. Math Biosci 168, 77–115 (2000).
56. Plank, M. J., Sleeman, B. D. & Jones, P. F. A mathematical model of tumour angiogenesis, regulated by vascular endothelial growth
factor and the angiopoietins. J Theor Biol 229, 435–454 (2004).
57. Stokes, C. L. & Lauffenburger, D. A. Analysis of the roles of microvessel endothelial cell random motility and chemotaxis in
angiogenesis. J Theor Biol 152, 377–403 (1991).
58. Anderson, A. R. A. & Chaplain, M. A. J. Continuous and discrete mathematical models of tumor-induced angiogenesis. Bull Math
59. Macklin, P. et al. Multiscale modelling and nonlinear simulation of vascular tumour growth. J Math Biol 58, 765–798 (2009).
60. Perfahl, H. et al. Multiscale modelling of vascular tumour growth in 3D: the roles of domain size and boundary conditions. PloS
Biol 60, 857–899 (1998).
ONE 6, e14790 (2011).
61. McDougall, S. R., Watson, M. G., Devlin, A. H., Mitchell, C. A. & Chaplain, M. A. J. A hybrid discrete-continuum mathematical
model of pattern prediction in the developing retinal vasculature. Bull Math Biol 74, 2272–2314 (2012).
62. Bauer, A. L., Jackson, T. L. & Jiang, Y. A cell-based model exhibiting branching and anastomosis during tumor-induced angiogenesis.
Biophys J 92, 3105–3121 (2007).
growth. PLoS Comp Biol 4, e1000163 (2008).
63. Merks, R. M., Perryn, E. D., Shirinifard, A. & Glazier, J. A. Contact-inhibited chemotaxis in de novo and sprouting blood-vessel
64. Shirinifard, A. et al. 3D multi-cell simulation of tumor growth and angiogenesis. PLoS ONE 4, e7190 (2009).
65. Perfahl, H. et al. 3D hybrid modelling of vascular network formation. Journal of theoretical biology 414, 254–268 (2017).
66. Milde, F., Lauw, S., Koumoutsakos, P. & Iruela-Arispe, M. L. The mouse retina in 3D: quantification of vascular growth and
remodeling. Integrative Biology 5, 1426–1438 (2013).
67. Travasso, R. D. M. The mechanics of blood vessel growth. In Vasculogenesis and Angiogenesis-from Embryonic Development to
Regenerative Medicine. InTechOpen, Rijeka Croatia, 187–204 (2011)
68. Mantzaris, N., Webb, S. & Othmer, H. Mathematical modeling of tumor-induced angiogenesis. J Math Biol 49, 111–187 (2004).
69. Spill, F., Guerrero, P., Alarcón, T., Maini, P. K. & Byrne, H. M. Mesoscopic and continuum modelling of angiogenesis. J Math Biol 70,
485–532 (2014).
70. Quinas-Guerra, M. M., Ribeiro-Rodrigues, T. M., Rodríguez-Manzaneque, J. C., Travasso, R. D. M. Understanding the dynamics of
tumor angiogenesis: A systems biology approach. in Systems Biology In Cancer Research and Drug Discovery (pp. 197–227). Springer
Netherlands (2012)
71. Heck, T., Vaeyens, M. M. & Van Oosterwyck, H. Computational models of sprouting angiogenesis and cell migration: towards
multiscale mechanochemical models of angiogenesis. Math Model Nat Phen 10, 108–141 (2015).
72. Guerra, M. M. D. S. Q. & Travasso, R. D. M. Novel approach to vascular network modeling in 3d. 2012 IEEE 2nd Portuguese Meeting
in Bioengineering (ENBENG), Feb, pp. 1–6 (2012)
73. Vilanova, G., Colominas, I. & Gomez, H. Capillary networks in tumor angiogenesis: From discrete endothelial cells to phase-field
averaged descriptions via isogeometric analysis. International Journal for Numerical Methods in Biomedical Engineering 29,
1015–1037 (2013).
74. Vilanova, G., Colominas, I. & Gomez, H. Coupling of discrete random walks and continuous modeling for three-dimensional
tumor-induced angiogenesis. Computational Mechanics 53, 449–464 (2014).
75. Emmerich, H. Advances of and by phase-field modelling in condensed-matter physics. Advances in Physics 57, 1–87 (2008).
76. Travasso, R. D. M., Castro, M. & Oliveira, J. C. R. E. The phase-field model in tumor growth. Philosophical Magazine 91, 183–206
77. Nonomura, M. Study on multicellular systems using a phase field model. PLoS ONE 7, e33501 (2012).
78. Boareto, M., Jolly, M. K., Ben-Jacob, E. & Onuchic, J. N. Jagged mediates differences in normal and tumor angiogenesis by affecting
tip-stalk fate decision. Proceedings of the National Academy of Sciences 112, E3836–E3844 (2015).
79. Benedito, R. et al. The notch ligands Dll4 and Jagged1 have opposing effects on angiogenesis. Cell 137, 1124–1135 (2009).
80. Palágyi, K. & Kuba, A. A 3D 6-subiteration thinning algorithm for extracting medial lines. Pattern Recognition Letters 19, 613–627
(2011).
(1998).
81. Vavourakis, V. et al. A validated multiscale in-silico model for mechano-sensitive tumour angiogenesis and growth. PLoS Comput
82. Grote, J. Tissue respiration. Human physiology, pp. 508–520, Springer (1983)
83. Maggelakis, S. & Savakis, A. A mathematical model of growth factor induced capillary growth in the retina. Mathematical and
Biol 13, e1005259 (2017).
computer modelling 24, 33–41 (1996).
84. Aubert, M., Chaplain, M. A., McDougall, S., Devlin, A. & Mitchell, C. A continuum mathematical model of the developing murine
retinal vasculature. Bulletin of mathematical biology 73, 2430–2451 (2011).
85. Popel, A. S. Theory of oxygen transport to tissue. Critical reviews in biomedical engineering 17, 257–321 (1988).
11
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8 86. Secomb, T. W., Hsu, R., Park, E. Y. & Dewhirst, M. W. Green's function methods for analysis of oxygen delivery to tissue by
microvascular network. s. Annals of biomedical engineering 32, 1519–1529 (2004).
87. Goldman, D. Theoretical models of microvascular oxygen transport to tissue. Microcirculation 15, 795–811 (2008).
88. Norton, K.-A. & Popel, A. S. Effects of endothelial cell proliferation and migration rates in a computational model of sprouting
angiogenesis. Scientific Reports 6, 36992 (2016).
89. Ulyatt, C., Walker, J. & Ponnambalam, S. Hypoxia differentially regulates VEGFR1 and VEGFR2 levels and alters intracellular
signaling and cell migration in endothelial cells. Biochemical and biophysical research communications 404, 774–779 (2011).
90. Nevo, O., Lee, D. K. & Caniggia, I. Attenuation of VEGFR-2 expression by sFlt-1 and low oxygen in human placenta. PloS one 8,
e81176 (2013).
91. Treps, L., Conradi, L.-C., Harjes, U. & Carmeliet, P. Manipulating angiogenesis by targeting endothelial metabolism: Hitting the
engine rather than the drivers–a new perspective? Pharmacological Reviews 68, 872–887 (2016).
92. Cantelmo, A. R., Pircher, A., Kalucka, J. & Carmeliet, P. Vessel pruning or healing: endothelial metabolism as a novel target? Expert
Opinion on Therapeutic Targets 21, 239–247 (2017).
93. Cantelmo, A. R. et al. Inhibition of the glycolytic activator PFKFB3 in endothelium induces tumor vessel normalization, impairs
metastasis, and improves chemotherapy. Cancer Cell 30, 968–985 (2016).
94. Schoors, S. et al. Partial and transient reduction of glycolysis by PFKFB3 blockade reduces pathological angiogenesis. Cell
metabolism 19, 37–48 (2014).
95. Ruhrberg, C. et al. Spatially restricted patterning cues provided by heparin-binding VEGF-A control blood vessel branching
morphogenesis. Genes & Development 16, 2684–2698 (2002).
96. Conradi, L. C. et al. Tumor vessel disintegration by maximum tolerable PFKFB3 blockade. Angiogenesis 20, 599–613 (2017).
97. Wenes, M. et al. Macrophage Metabolism Controls Tumor Blood Vessel Morphogenesis and Metastasis. Cell Metabolism 24, 701–715
(2016).
Acknowledgements
The authors thank fruitful discussions with Lino Ferreira and Paulo Matafome. The simulations were done in the
Navigator cluster of the University of Coimbra. This work was funded by FEDER funds through the Operational
Programme Competitiveness Factors - COMPETE and by national funds by FCT - Foundation for Science and
Technology under the strategic project UID/FIS/04564/2016 (M.M.S., J.C., R.D.M.T.). M.M.S. acknowledges
the support of the National Council of Technological and Scientific Development (CNPq) under the grant
235101/2014-1. RT acknowledges FCT's support through the FCT Researcher Program.
Author Contributions
M.M.-S., J.C. and R.D.M.T. conceived the study; M.M.-S. and R.D.M.T. wrote the code; M.M.-S., R.C. and L.R.
run the simulations; M.M.-S., J.C. and R.D.M.T. wrote the manuscript.
Additional Information
Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-018-27034-8.
Competing Interests: The authors declare no competing interests.
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. The images or other third party material in this
article are included in the article's Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article's Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© The Author(s) 2018
12
www.nature.com/scientificreports/SCIEnTIFIC REPORTS (2018) 8:8726 DOI:10.1038/s41598-018-27034-8 |
1109.1155 | 3 | 1109 | 2011-11-14T16:57:07 | Crowding of molecular motors determines microtubule depolymerization | [
"physics.bio-ph",
"cond-mat.stat-mech",
"q-bio.SC"
] | Assembly and disassembly dynamics of microtubules (MTs) is tightly controlled by MT associated proteins. Here, we investigate how plus-end-directed depolymerases of the kinesin-8 family regulate MT depolymerization dynamics. Employing an individual-based model, we reproduce experimental findings. Moreover, crowding is identified as the key regulatory mechanism of depolymerization dynamics. Our analysis gives two qualitatively distinct regimes. For motor densities above a particular threshold, a macroscopic traffic jam emerges at the plus-end and the MT dynamics become independent of the motor concentration. Below this threshold, microscopic traffic jams at the tip arise which cancel out the effect of the depolymerization kinetics such that the depolymerization speed is solely determined by the motor density. Because this density changes over the MT length, length-dependent regulation is possible. Remarkably, motor cooperativity does not affect the depolymerization speed but only the end-residence time of depolymerases. | physics.bio-ph | physics |
Crowding of molecular motors determines
microtubule depolymerization
Louis Reese1
Anna Melbinger2
Erwin Frey3
Arnold Sommerfeld Center for Theoretical Physics and
Center for NanoScience, Department of Physics,
Ludwig-Maximilians-Universität München,
Theresienstrasse 37, D-80333 Munich, Germany
October 31, 2018
1Email: [email protected]
2Email: [email protected]
3Corresponding author. Email: [email protected]
Abstract
Assembly and disassembly dynamics of microtubules (MTs) is tightly controlled by
MT associated proteins. Here, we investigate how plus-end-directed depolymerases
of the kinesin-8 family regulate MT depolymerization dynamics. Employing an
individual-based model, we reproduce experimental findings. Moreover, crowding
is identified as the key regulatory mechanism of depolymerization dynamics. Our
analysis gives two qualitatively distinct regimes. For motor densities above a par-
ticular threshold, a macroscopic traffic jam emerges at the plus-end and the MT
dynamics become independent of the motor concentration. Below this threshold,
microscopic traffic jams at the tip arise which cancel out the effect of the de-
polymerization kinetics such that the depolymerization speed is solely determined
by the motor density. Because this density changes over the MT length, length-
dependent regulation is possible. Remarkably, motor cooperativity does not affect
the depolymerization speed but only the end-residence time of depolymerases.
Key words: kinesin-8; length-regulation; microtubule dynamics; traffic jam;
particle exclusion; driven transport
INTRODUCTION
Microtubules (MTs) are cytoskeletal filaments that serve a central role in intra-
cellular organization (1, 2) and several cellular processes including mitosis (3, 4),
cytokinesis (5) and intracellular transport (6). They can cope with this multitude
of diverse tasks because they are highly dynamic structures which continually as-
semble and disassemble through the addition and removal of tubulin heterodimers
at their ends. GTP-hydrolysis is the energy source which drives switching between
persistent states of growth and shrinkage, a stochastic process termed dynamic in-
stability (7 -- 10). Each cellular process employs a specific set of MT-associated
proteins (MAPs) to tightly regulate the rates of growth and shrinkage as well as
the rate of transition between these states (11 -- 13).
Depolymerases from the kinesin-8 and kinesin-13 protein families (e.g., Kip3p
and MCAK, respectively) are important regulators of MT dynamics. They are
thought to promote switching of MTs from growth to shrinkage (catastrophes)
(12). Whereas MCAK lacks directed motility and diffuses along MTs (14), Kip3p
is a highly processive plus-end-directed motor (15, 16). Proteins from the kinesin-8
family are important for regulating MT dynamics in diverse organisms. Kif18A
is a key component in chromosome positioning in mammalian cells (17 -- 19) where
it regulates plus-end dynamics. Its orthologs, the plus-end directed motors Kip3p
in budding yeast (16) and Klp5/6 in fission yeast (20 -- 22), show depolymerizing
activity. A notable feature shared by these MT plus-end depolymerases is that they
depolymerize longer MTs more rapidly than they do shorter ones (15, 17, 21, 23).
A similar length-dependent regulation of MT assembly by kinesin-5 motors was
observed in in vivo studies of chromosome congression in budding yeast (24). The
key experimental observations from in vitro studies of Kip3p (23) are that 1), the
end-residence time of Kip3p at the tip depends on the bulk concentration of Kip3p
and correlates inversely with the macroscopic depolymerization speed; and 2), the
macroscopic depolymerization rate is directly proportional to the flux of Kip3p
towards the MT plus-end.
It is thought that length-dependent depolymerization kinetics serves several
purposes (2). For example, positioning of the nucleus at the cell center during
interphase is achieved by growing MTs that push against the cell poles while re-
maining attached to the nucleus. A higher rate of catastrophes for longer MTs
implies that shorter MTs have an increased contact time with the cell poles. Com-
puter simulations show that this leads to a higher efficiency of nuclear positioning
during interphase (25).
There is convincing experimental evidence that molecular traffic along MTs
strongly affects MT depolymerization dynamics. However, in vitro experiments
can not yet fully explore the underlying traffic dynamics. Theoretical investigations
employing individual-based models can be instrumental in furthering a mechanistic
2
understanding of this process. Fortunately, these models can be constructed on the
basis of substantial quantitative data available from in vitro experiments (15, 23)
characterizing the binding kinetics and the motor activity of plus-end-directed
motors. Therefore we sought to identify the molecular mechanisms underlying
the observed correlation between depolymerization dynamics and molecular traffic
along MTs.
In this study, we constructed an individual-based model for the coupled dynam-
ics of MT depolymerization and molecular traffic of plus-end-directed motors. This
model quantitatively reproduces previous experimental results (15, 23). Moreover,
we make precise quantitative predictions for the density profiles of molecular mo-
tors on the MT and demonstrate that molecular crowding and ensuing traffic jams
regulate the depolymerization dynamics. We find two qualitatively distinct regimes
of depolymerization dynamics: At low bulk concentrations of depolymerases, the
depolymerization speed of MTs is density-limited and is a function of the bulk
concentration and average motor speed alone. There is a sharp threshold in bulk
depolymerase concentration above which macroscopic traffic jams emerge and the
depolymerization speed is simply given by the microscopic depolymerization rate.
Of note, none of these features are affected by the degree of cooperativity in the
depolymerization kinetics. In contrast, the end-residence time of a depolymerase
(i.e., the typical time it spends at the plus-end) is strongly correlated with coop-
erativity. We outline how these predictions from our theoretical analysis can be
tested experimentally.
RESULTS
Model definition
We use an individual-based model, as illustrated in Fig. 1, to describe the dy-
namics of plus-end-directed depolymerases. Motor proteins, present at a constant
bulk concentration c, are assumed to randomly bind to and unbind from the MT
lattice with rates ωa and ωd, respectively. Bound motors are described as Pois-
son steppers (A more detailed biochemical model for motors on microtubules has
to await further experimental analysis. One of the different possible schemes has
recently been studied by Klumpp et al. (26).) that processively walk along individ-
ual protofilaments towards the plus-end at an average speed v (27). These motors
hinder each other sterically because individual binding sites i = 1, . . . , L on each
protofilament can either be empty (ni = 0) or occupied by a single motor (ni = 1).
Because switching between protofilaments is rare (27), transport along each of
the protofilaments can be taken as independent, and the model becomes effec-
tively one-dimensional (28) (Fig. 1B). Models of this type were recently discussed
3
FIGURE 1: Illustration of MT and motor dynamics. Molecular motors present at
concentration c randomly attach to unoccupied tubulin dimers along the MT lat-
tice with rate ωa. While bound they processively move toward the plus-end at rate
ν, and unbind with rate ωd. Because motors do not switch lanes (protofilaments),
the MT lattice (A) becomes effectively one-dimensional (B). Each lattice site ni
(with i = 1, . . . , L numbering the sites) may be empty (ni = 0) or occupied by a
single motor (ni = 1). At the plus-end the motors act as depolymerases (indicated
by scissors) either alone with rate δ0 or cooperatively with rate δ1.
as minimal models for intracellular transport (29 -- 32). In its given formulation,
where the cytosol is considered as a homogeneous and constant reservoir of motors,
it is equivalent to a driven lattice gas model known as the totally asymmetric sim-
ple exclusion process with Langmuir kinetics (TASEP/LK) (29). A central finding
of this model is that the interplay between on-off (Langmuir) kinetics and directed
transport along protofilaments can result in "traffic jams" in which the density pro-
file of motors along a protofilament shows a sharp increase from a low-density to a
crowded high-density regime (29, 31). Such and other crowding effects (33, 34) are
important for a molecular understanding of MT dynamics. Previous theoretical
studies on this topic largely disregarded crowding effects or considered parame-
ter regimes in which they are unimportant (35 -- 37). Depolymerization, including
crowding effects, has also been investigated for diffusive depolymerases such as
MCAK (38).
4
ωdδ0δ1ABνccωanon-cooperativecooperativeAt the plus-end of the system we consider depolymerization dynamics arising
due to the interaction of molecular motors with the MT tip. Motivated by recent
experiments (23), we assume non-processive depolymerization, i.e, a molecular
motor dissociates from the lattice after triggering depolymerization. Because the
molecular mechanisms are not yet fully resolved, we study two scenarios of de-
polymerization (see Fig. 1B). In the noncooperative scenario, the dissociation rate
depends only on whether the last site is occupied by a motor. If the last site is oc-
cupied, nL = 1, the MT depolymerizes at rate δ0. However, recent single molecule
studies indicate that Kip3p may act cooperatively (23), which we consider as our
second scenario. After arriving at the plus-end, the motor is observed to pause
and depolymerize a tubulin dimer only after a second Kip3p has arrived behind
it. In this scenario, a tubulin dimer is depolymerized with rate δ1 if both the last
and the second-to-last sites are occupied, nL−1 = nL = 1. Therefore, the total
depolymerization rate can be written as:
∆ = δ0nL + δ1nL−1nL .
(1)
For stabilized MTs, the spontaneous depolymerization rate is small (23) and thus is
not considered here. The relative magnitude of the noncooperative rate δ0 and the
cooperative rate δ1 determines the degree of cooperativity of the depolymerization
kinetics. In an average over many realizations of the stochastic process (ensemble
average), the depolymerization speed Vdepol depends on the occupation of the last
two binding sites by depolymerases (Fig. 1B):
Vdepol = (δ0ρ+ + δ1κ+) a ,
(2)
where a is the lattice spacing. Here ρ+ := (cid:104)nL(cid:105) is the probability that the last site
is occupied (i.e., the expected motor density at the plus-end), and κ+ := (cid:104)nL−1nL(cid:105)
denotes the probability that both the last and second-to-last sites are occupied.
We analyzed this model via stochastic simulations and analytic calculations (for
further details, see the Supporting Material).
Validation of the model and its parameters
The model parameters are, as far as they are available, fixed by experimental data.
The motor speed, v, the motor run length, (cid:96), and motor association rate, ωa, were
measured previously (23):
v = 3.2 µm min−1 ,
ωa = 24 nM−1min−1µm−1 ,
(cid:96) ≈ 11 µm .
5
Using an MT lattice spacing of a = 8.4 nm, we derive the corresponding parameters
in our model as follows: The motor speed v corresponds to 6.35 lattice sites per
second, i.e., a hopping rate of ν = v/a = 6.35 s−1. The inverse hopping rate
τ := ν−1 = 0.16 s and the size a of a tubulin dimer serve as our basic timescale
and length scale, respectively. Then, the measured association rate corresponds
to a rate ωa ≈ 5.3 × 10−4 nM−1site−1 τ−1. The dissociation rate, ωd = v/(cid:96), is
derived as the ratio of the mean motor speed, v, and the mean motor run-length,
(cid:96). The latter equals 1310 lattice sites. Thus, the dissociation rate is expressed
as ωd ≈ 7.6 × 10−4 site−1 τ−1. In contrast to the transport behavior on the MT,
the parameters concerning the depolymerization rates, δ0/1, cannot be directly
extracted from experiments. However, there is evidence for a depolymerization
rate as high as the motor speed, v (15, 23). As a starting point for the following
discussion we tentatively take δ0 = ν.
Using the above set of parameters we now phenomenologically compare the
results from numerical simulations of our model with observations from experi-
ments. Specifically, we consider kymographs of the MT, which show how the MT
length and the motor density on the MT evolve over time. For the simulation data
shown in Fig. 2 we consider an MT consisting of 14 independent protofilaments
and investigate the dynamics for the noncooperative scenario and a range of motor
concentrations, c = 1.2, 1.8, 2.6 nM, cf. Fig. 2A-C. Surprisingly, as shown later,
neither the cooperativity of the motors nor a decrease in the depolymerization
rates led to different shapes of kymographs (see also Fig. S1). We find an initial
time period in which, starting from an empty MT lattice, the motors first fill up
the lattice (39, 40). This is followed by a time window in which the motor den-
sity exhibits a quasi-stationary profile, i.e., the density at a certain distance from
the minus-end does not change except for boundary effects induced by the plus-
end. The corresponding density profiles are illustrated in Fig. 2E and discussed
in more detail in the following section. In this quasi-stationary regime, the de-
polymerization dynamics shows qualitatively different behavior depending on the
concentration of free motor molecules: At low concentration, c < 1.4 nM, and thus
low density of motors on the MT, depolymerization slows down gradually in the
course of time (Fig. 2A). When the motor concentration increases to larger values,
c > 1.4 nM, an intermediate regime emerges in which the depolymerization speed
stays roughly constant (Fig. 2B and C). Remarkably, we find that during this
regime, the depolymerization speed is directly proportional to the motor density,
Vdepol(L) = ρ−(L)v (Fig. 2D). At a third stage in the depolymerization process,
there is a rather abrupt change in the depolymerization speed right where the
density profile shows a steep drop (Fig. 2C-E). After we have elaborated more on
the theoretical model, we discuss why there is such a tight correlation between the
depolymerization dynamics and the density profile.
6
FIGURE 2: Validation of the theoretical model.
(A)-(C) Time-space plots of
stochastic simulations for a range of motor concentrations and depolymerization
rate δ0 = 6.35 sites s−1. The density of molecular motors is shown as the bright
area (green), and the MT is shown as the dim area (red; for details, see Supporting
Material). For low concentrations, c < 1.4 nM, depolymerization slows down grad-
ually (23). At higher concentrations, c > 1.4 nM, there is a rather abrupt change in
MT shortening. This change is correlated with a steep decrease in the motor den-
sity (DW), indicated as dotted lines. (D) The depolymerization speed, Vdepol, as a
function of the length of the MT L(t), extracted from the simulation data shown
in the kymograph (gray). The position of the DW (dotted), and the predicted
depolymerization speed, Vdepol = vρ(L) (see also Eq. 10), using the linear approxi-
mation for the motor density profile (black) and the density profile extracted from
stochastic simulations (green), coincide very well with the observed depolymer-
ization speed; v = 6.35 sites s−1 is the walking speed of the motors. (E) Density
profiles at the minus-end from stochastic simulations (lines with symbols), exact
solutions (solid) and linearized theory (dotted) are shown. (F) As a function of
the motor concentration, c, and the distance from the minus-end, there are dis-
tinct types of density profiles. At motor concentration lower than c = 1.4 nM
(thin black), the density of motors along the MT is low and the profile is smooth.
The Langmuir density is reached continuously after a certain MT length (dashed,
numerical). At high concentrations, c > 1.4 nM there are two regions along the
MT separated by an intervening DW (black, exact; see Supporting Material): an
approximately linear antenna profile and a flat profile (Langmuir density). Linear
approximations for the continuous and the discontinuous transitions (Eq. 4) are
shown as well (gray). Thin lines refer to the density profiles shown in E.
7
ABCDEFAntenna ProfileLangmuir DensityDomain WallContinuousTransition(Wedge)All of these qualitative features of MT dynamics are identical to those found ex-
perimentally (15, 23), and suggest that the density profile and, in particular, traffic
jams formed on the MT lattice are the main determinants of the depolymerization
dynamics. Moreover, the time scales of the dynamics agree quantitatively well
with experimental results for the same motor concentrations (15, 23). This vali-
dates our theoretical model because up to the depolymerization rate δ, all of the
model parameters were derived from experimental data (23).
Density profiles at the minus-end (bulk density)
The above observations strongly point toward a tight correlation between the de-
polymerization speed and the motor density profile at the minus end, ρ−(x), which
we henceforth call the bulk (motor) density. The quasi-stationary bulk density pro-
files shown in Fig. 2E were obtained by assuming very long lattices; effects caused
by the plus-end are not visible in the vicinity of the minus-end. A more detailed
discussion of these simulations can be found in the Supporting Material. Since this
bulk density will play an important role in the following analysis, we summarize
its features as obtained from analytical calculations detailed in the Supporting
Material.
At the minus-end, the density profiles show an initial linear increase. This is
an "antenna effect" (15) as illustrated in Fig. 3A. Motors that attach in proximity
of the MT minus-end immediately move toward the plus-end, thereby generating
an approximately linearly increasing accumulation of motors. The slope is given
by K/(cid:96), where K = c ωa/ωd denotes the binding constant. At sufficiently large
distances from the minus-end, the density profile becomes flat and dominated by
Langmuir kinetics with the ensuing Langmuir density:
ρLa =
K
1 + K
=
c ωa
c ωa + ωd
.
(3)
The full density profile is obtained by concatenating the antenna profile and the
flat Langmuir profile such that the motor current is continuous along the MT.
We find two qualitatively distinct scenarios (Fig. 2E). For low concentrations of
molecular motors, c, the antenna profile matches the asymptotic Langmuir density
continuously, resulting in a wedge-like profile. In contrast, above a certain thresh-
old value for the concentration, determined by the binding constant K−
c = 1, the
two profiles can no longer be matched continuously and the density profile displays
a sharp discontinuity, also termed a "domain wall" (DW) (29). In other words, if
the Langmuir density rises above a critical value of ρc
La = 0.5, a crowding-induced
traffic jam will result (41) (Fig. 3A). The density profiles obtained from the an-
alytic calculations and the stochastic simulations agree nicely, as illustrated in
8
FIGURE 3: Illustration of the antenna and crowding regimes, and of cooperativity.
(A) Starting from an empty MT, motors start to accumulate on the MT lattice
by attachment and subsequent transport to the plus-end. The combined effect of
Langmuir kinetics and steric exclusion between the motors leads to two sharply
separated regimes. Starting from the minus-end, the motor density increases lin-
early (antenna profile). At a certain critical length (cid:96)−, a macroscopic traffic jam
arises because particles hinder each other and crowding dominates the MT density.
(B) Illustration of non-cooperative (B, nc) and fully cooperative (C, fc) depoly-
merization kinetics. With regard to the the depolymerization speed, both models
are effectively equal (see main text).
Fig. 2E. In particular, the theoretical analysis gives an explicit expression for the
width of the antenna-like profile:
(cid:40) 1
(cid:96)− ≈ (cid:96)
1+K
1
K(1+K)
for K < 1 ,
for K > 1 .
(4)
This result reduces to the average run length of molecular motors, (cid:96) = v/ωd,
in the limit of very low binding constant, K (cid:28) 1, where crowding effects can be
neglected (37). However, with increasing K the regime with an antenna-like profile
becomes significantly shorter than (cid:96) (Fig. 2F).
9
BCCrowdingAntennaTransportMinus-EndPlus-End−A(Langmuir)Depolymerization dynamics is independent of cooperativity
We now address how the cooperativity of the depolymerization kinetics affects the
macroscopic depolymerization speed. There are two limiting cases: noncooperative
depolymerization (nc) with (δ0, δ1) = (δ, 0), and fully cooperative depolymeriza-
tion (fc) with (δ0, δ1) = (0, δ) (for an illustration, see Fig. 3, B and C). Remarkably,
we find from our stochastic simulations, shown in Fig. 4, that there is no difference
in depolymerization speed for these two limiting cases. Even when the depolymer-
ization dynamics contains cooperative as well as noncooperative terms, we do not
find any significant differences in the depolymerization speed (Fig. 4B).
This observation from our stochastic simulations can be explained by the follow-
ing molecular mechanism: Consider a model with fully cooperative depolymeriza-
tion kinetics. Then, after the first motor has arrived at the plus-end, the terminal
site of the MT will remain occupied from that time on. Depolymerization only
occurs if another motor arrives at the second-to-last site. In other words, while the
last site remains occupied, the second-to-last site triggers the depolymerization.
Hence, as far as the depolymerization speed is concerned, the fully cooperative
model is identical to a noncooperative model with the same molecular rate δ. In
the noncooperative model the terminal tubulin dimer is removed at rate δ once a
molecular motor has arrived at the last site (see Fig. 3B). In the fully cooperative
model, the terminal tubulin dimer is removed once a molecular motor has arrived
at the second-to-last site next to a permanently occupied last site (Fig. 3C).
Depolymerization dynamics is strongly affected by crowding
To gain further insights in the correlation between the depolymerization speed and
the density of motors on the MT, we performed stochastic simulations focusing
on the MT plus-end by regarding the dynamics in a co-moving frame.
Instead
of simulating the full-length MT with an antenna profile and a subsequent flat
Langmuir density, we considered a reduced model in which the density at the left
end is set equal to the Langmuir density ρLa. For long MTs, the Langmuir density
is always reached, so that the reduced system is fully equivalent to the original
model. Our simulations show two clearly distinct regimes of depolymerization
dynamics (Fig. 4): For small microscopic depolymerization rates, δτ < ρLa, the
depolymerization speed is rate-limited: Vdepol = aδ. In contrast, for rates δτ >
ρLa, the depolymerization speed is density-limited, and the Langmuir density is
the limiting factor: Vdepol = ρLav. The boundary between the two regimes is
remarkably sharp and given by
ρ∗
La = δτ .
(5)
This implies that the depolymerization speed can switch between being density-
limited and rate-limited by changing the concentration c or the values of the bio-
10
chemical rates of depolymerases binding to and unbinding from the MT lattice.
Overall, the depolymerization speed obeys a scaling law
Vdepol = ρLav V(δτ /ρLa) =
aδ
ρLav
for δτ ≤ ρLa
for δτ > ρLa
,
(6)
(cid:40)
where V(x) is a universal scaling function with the simple form V(x) = x for x < 1
and V(x) = 1 for x > 1. Experimentally, this implies that one should find data
collapse upon using such a scaling plot (Fig. 4A).
To gain a molecular understanding of these remarkable features of the depoly-
merization speed, one needs to have a closer look at the density profile of the
molecular motors at the MT tip.
If the depolymerization rate is small, δ < ν,
motors leave the tip more slowly than they arrive. Therefore, the MT tip acts as a
bottleneck for molecular transport that disturbs the density profiles either locally
or macroscopically. A weak bottleneck induces a local perturbation ("spike") (33).
These spikes are sharp changes of the density profile with a typical extension that
scales with the size of a heterodimer. However, if the strength of a bottleneck
exceeds a threshold value, the spike extends to a macroscopic perturbation ("traffic
jam") (33). Fig. 5A illustrates how, for a given Langmuir density, ρLa = 2/3, the
effect on the density profile changes from a spike (blue) to an extended traffic jam
(red and green) when the depolymerization rate is δ.
Let us now analyze the conditions and consequences of such bottlenecks in
more detail. Suppose we are in a parameter regime where the plus-end disturbs
the density profile only locally, i.e., on the scale of a heterodimer. Then, we may
take the bulk density to be equal to the Langmuir density, ρLa, up to the last site
(the plus-end) where it jumps to some higher or lower value ρ+. The particle loss
current at the plus-end due to MT depolymerization is then given by
Jdepol = (1 − ρLa)ρ+δ .
(7)
The factor 1 − ρLa arises because the particle number decreases only if a particle
depolymerizes the MT and the second-to-last site, L−1, is unoccupied. Otherwise,
depolymerization dynamics and the associated frame shift of the MT lattice do
not change the occupation of the last site. This particle loss has to be balanced
by the incoming particle flux,
JLa = ρLa(1 − ρLa)ν .
(8)
Equating these particle fluxes (Eqs. 7 and 8) implies the following condition for
the motor density at the plus-end:
for ρLa ≤ δτ
for ρLa > δτ
ρLa/δτ
1
(cid:40)
(9)
ρ+ =
,
11
FIGURE 4: Scaling plot for the depolymerization speed Vdepol. (A) Upon rescaling,
both the macroscopic depolymerization speed, Vdepol, and the microscopic depoly-
merization rate, δ, with the Langmuir density, ρLa, all data collapse onto one uni-
versal scaling function V (solid gray). A sharp transition at δτ = ρ∗
La distinguishes
the rate-limited regime from the density-limited regime. (B) Comparison of co-
operative and noncooperative depolymerization, with the macroscopic depolymer-
ization speed Vdepol as a function of Langmuir density ρLa. For δ := δ0 + δ1 = 0.7 ν
different degrees of cooperativity are displayed as indicated in the graph.
12
ABFIGURE 5: Density profiles at the plus-end, corresponding phase diagram, and
depolymerization scenarios. (A) Density profiles at the MT plus-end in the co-
moving frame for c = 2.9 nM, and δ = 0.1, 0.3 (left), 0.35, 0.5 (middle) and 0.8 ν
(right). The simulation results and analytical solutions (black; see Supporting Ma-
terial) agree nicely. (B) Depending on the value of δ and the density of motors,
ρLa, there are three different classes of density profiles at the plus-end: wedge-like
(diamonds), traffic jams with a DW (square), and spikes (circles). The transition
between profiles with an extended traffic jam and a localized spike (solid line) also
marks a qualitative change in the depolymerization speed. Whereas the depoly-
merization speed is density-limited in the spike regime, it is rate-limited in the
DW and wedge regime. Symbols correspond to parameters as displayed in panel
A. (C) Depending on the value of δ and the density of motors, ρLa, there are three
different regimes of depolymerization dynamics. In regime α depolymerization is
density-limited for arbitrary MT length.
In contrast, depolymerization is rate-
limited for long MTs and density limited for short MTs in the regimes β and γ.
For details, see the main text.
13
CABwhere the fact that the motor density is bounded ρ+ ≤ 1 is already accounted
for. The particle density on the last site, in turn, determines the depolymerization
speed. For ρLa < δτ, one obtains according to Eq. 2 and Eq. 9:
Vdepol = ρ+ δa = ρLav .
(10)
Remarkably, here the effect of the depolymerization kinetics (δ) cancels out such
that the macroscopic depolymerization speed is independent of the molecular de-
tails of depolymerization kinetics and solely determined by the Langmuir density,
i.e., the motor density in the bulk, ρ−(x), and not at the tip of the MT. This
result crucially depends on the presence of a microscopic spike. It explains the
hitherto puzzling experimental result that the depolymerization speed is directly
proportional to the bulk motor current along the MT (23) (Fig. S2).
Because the density is bounded, ρ+ ≤ 1, density profiles with a spike are
only possible if the densities are not too large, ρLa < δτ. This is the case for
the blue curve in Fig. 5A. For densities exceeding the critical density, ρ∗
La = δτ,
the bottleneck-induced perturbation in the density profile can no longer remain
a local spike, but has to become macroscopic in extent (33) (see green and red
curves in Fig. 5A and Supporting Material). One finds that over an extended
region, the binding sites at the plus-end then remain permanently occupied such
that ρ+ = 1. This immediately implies that the depolymerization speed becomes
density-independent and proportional to the microscopic depolymerization rate:
Vdepol = aδ .
(11)
There is a tight correlation between the shape of the density profiles and the
macroscopic depolymerization speed. The analytic results explain the molecular
mechanism behind the numerically observed scaling law (Eq. 6), with a sharp
transition from density-regulated to a rate-limited depolymerization dynamics at
a critical value of ρ∗
La = δτ (cf. the classification of density profiles and depoly-
merization regimes shown in Fig. 5B).
Actually, the above calculations can be generalized to the regime in which the
motor density exhibits an antenna-like linear profile, i.e., for MT length shorter
than (cid:96)−. As detailed in the Supporting Material, we find that the depolymerization
speed is rate-limited, Vdepol = aδ, if MTs are shorter than (cid:96)− but still longer than
a second threshold length:
(cid:96)d := δa/cωa = (cid:96) δτ /K .
(12)
In contrast, for (cid:96)d > (cid:96)−, the depolymerization speed in the antenna regime is always
length-dependent and strictly follows the shape of the antenna profile, ρ−(x):
Vdepol = ρ−(L)v .
(13)
14
Using Eq. 4, the condition (cid:96)d > (cid:96)− on the threshold lengths is equivalent to
δτ > ρLa for K < 1 and to δτ > 1 − ρLa for K > 1.
Combining all of the above results, we find three mechanisms governing de-
polymerization dynamics, as illustrated in Fig. 5C:
(α) For δτ > ρLa, the depolymerization speed is always density-regulated and
given by Vdepol(L) = ρ−(L)v, where L is the time-dependent length of the
MT. In this parameter regime, the depolymerization speed is a direct map
of the bulk motor density profile on the MT, ρ−(x), a feature that can be
exploited experimentally to measure the profile.
(β) For ρLa > δτ > 1 − ρLa, the depolymerization speed is rate-limited for MTs
longer than (cid:96)−, and becomes density-limited as soon as the MT length falls
below (cid:96)− where the density profile is antenna-like. This implies that there is
a discontinuous jump in the depolymerization speed right at L = (cid:96)−.
(γ) Finally, for all other values for δτ, the depolymerization speed of the MT re-
mains rate-limited for lengths larger than a threshold length (cid:96)d. At (cid:96)d, which
is smaller than (cid:96)− in this parameter regime, there is again a discontinuous
jump to a density-limited depolymerization dynamics.
If the depolymerization rate is larger or equal to the hopping rate of molecular
motors, δτ ≥ 1, then δτ > ρLa is always obeyed simply because ρLa ≤ 1. In this
regime, all of the molecular details of the depolymerization kinetics are irrelevant.
Neither cooperativity nor the actual value of the depolymerization rate matters
in terms of the depolymerization speed; instead, only the bulk density regulates
the speed. Note that this was the case for the data shown in Fig. 2, where we
tentatively made the parameter choice δτ = 1. If the motors are faster than the
depolymerization process, δτ < 1, we have to distinguish between the parameter
regimes (α,β, and γ, Fig. 5C). Here the value of the depolymerization rate matters
if the bulk density exceeds a certain threshold concentration, ρLa > δτ, and the
MTs are long enough. Finally, the depolymerization speed always becomes density-
dependent and hence length-dependent if the MT length is short enough; the
corresponding threshold length is (cid:96)reg = min[(cid:96)−, (cid:96)d].
End-residence time strongly depends on cooperativity
In contrast to the depolymerization speed, the mean end-residence time τres is
strongly affected by the degree of cooperativity. Fig. 6 displays τres as obtained
from our stochastic simulations for noncooperative and fully cooperative depoly-
merization kinetics. Our simulations show that the end-residence time for the fully
cooperative model is identical to the average lifetime of a terminal tubulin dimer
15
τ fc
res = τd := a/Vdepol (Fig. 6A). Even for the noncooperative model, τ nc
res equals τd
for large residence times and deviates from it only at small values. The relatively
sharp transition to a constant lifetime of the terminal tubulin dimer occurs right
at τ nc
res = τ /ρLa, i.e., the end-residence time equals the waiting time for a molecu-
lar motor to arrive at the MT tip. For τ nc
res < τ /ρLa, the lifetime of the terminal
tubulin dimer is identical to the arrival time (Fig. 6,A and B). Once the arrival
time becomes shorter than the inverse depolymerization rate, the end-residence
time levels off at τ nc
res = 1/δ. These results show that the dependence of the end-
residence time on density can be used to quantify the degree of cooperativity.
This would require experiments with motor densities on the MT larger than those
studied up to now (15, 23).
The observation that the depolymerization speed is independent of the degree
of cooperativity seems to be at odds with the experimental finding that the end-
residence time, τres, of Kip3p depends on the total Kip3p concentration and is
inversely proportional to the macroscopic depolymerization speed (23). Actually,
however, there is no contradiction and the findings are readily explained within
our theoretical model: For a noncooperative model, τ nc
res is simply given by the
depolymerization rate, because after they arrive, the particles stay at the tip until
they depolymerize the MT:
τ nc
res =
1
δ
.
(14)
+
1
δ
ρLa
(cid:19)
For a fully cooperative model, τ fc
res depends not only on δ, but also on the rate
at which the second-to-last site becomes populated. Say the probability for the
second-to-last site to be occupied is ρ+. Then, τ fc
res is given by a sum of two
contributions arising from the cases in which the second-to-last site is empty or
occupied, respectively:
res = (1 − ρ+)
τ fc
(15)
If-the-second to last site is empty (which is the case with probability 1 − ρ+) τres
is the sum of arrival time τ /ρLa and depolymerization time 1/δ. Otherwise, the
end-residence time τres simply equals 1/δ.
(cid:18) τ
As shown in the previous section, two distinct scenarios arise: For small bulk
densities such that ρLa < δτ, the density profile at the plus-end exhibits a micro-
scopic spike with ρ+ = ρLa/δτ. For large densities, ρLa > δτ, a macroscopic traffic
jam emerges such that ρ+ = 1. This result obtained for the motor density at the
MT tip (Eq. 9) may now be used to calculate τ fc
res using Eq. 15:
1
δ
.
+ ρ+
(cid:40) 1
δ
τ
ρLa
τ fc
res =
for ρLa > δτ ,
else .
16
(16)
FIGURE 6: Motor end-residence times τres for cooperative and noncooperative
depolymerization.
(A) Mean end-residence time τres plotted against the mean
depolymerization time τd. Data were recorded for a range of depolymerization
rates δ = 0.02 . . . 2 ν. Noncooperative (shaded) and cooperative (black) dynamics
are shown for different densities. (B) Mean end-residence time τres as a function
of the Langmuir density ρLa for various depolymerization rates (in units of ν).
For noncooperative depolymerization, τres is given by 1/δ (shaded lines). For the
fully cooperative scenario (symbols), τres depends on whether the system is in the
density-limited (δτ > ρLa) or in the rate-limited (δτ < ρLa) regime. While, for
δτ > ρLa, the end-residence time is given by τres = τ /ρLa (solid gray line), for
δτ < ρLa, it is density-independent and determined by the microscopic depolymer-
ization rate τres = 1/δ (see also Eq. 16).
17
ABThis agrees well with the results from stochastic simulations displayed in Fig. 6. A
comparison with Eq. 6 shows that the end-residence time equals the typical depoly-
merization time, i.e., the expected lifetime of a terminal tubulin dimer, τ fc
res = τd.
This is in agreement with experimental findings regarding the unbinding-rate of
motors at the plus-end (23) and strongly supports the conclusion that depoly-
merization of MTs by Kip3p is fully cooperative. Varga et al. (23) measured the
end-residence time of motors on double stabilized MTs, i.e., where depolymer-
ization is switched off. They observed that the end-residence time is inversely
correlated with the concentration of Kip3p, and fit their data with an exponential
using a cut-off. This is in accordance with our results shown in Fig. 6B. However,
since depolymerization has been switched off in the experiment, the rate δ, corre-
sponding to the cutoff, now has to be interpreted as an unbinding-rate of motors
at the plus-end. It would be highly interesting to design experiments where the
depolymerization kinetics remains switched on, because this would allow one to
measure the magnitude of the microscopic depolymerization rate δ.
DISCUSSION
In this work, we have analyzed the effect of crowding and cooperativity on the
depolymerization dynamics of microtubules. To that end, we constructed an
individual-based model for the coupled dynamics of plus-end directed motor traffic
and microtubule depolymerization kinetics. The model is based on well-established
molecular properties of motors from the kinesin-8 family, i.e., the motors move on
single protofilaments with high processivity at an average speed v, and exchange
of motors between the bulk and the microtubule follows Langmuir kinetics. All
parameters of the model, including the average walking speed, run length, and at-
tachment rate, were directly extracted from available in vitro data (23). We have
validated our model by reproducing the onset of length-dependent depolymeriza-
tion as studied recently (15, 23). Without using any additional fitting parameter,
we found the same regimes of density profiles and ensuing depolymerization dy-
namics as in the experiments, i.e., a linear antenna-profile with a length-dependent
depolymerization speed and a flat profile with a constant depolymerization speed.
Moreover, we identified a threshold density of motors above which a crowding-
induced traffic jam emerges at the minus-end. The predicted shape and extent of
these traffic jams should be amenable to experiments that raise the depolymerase
concentration c or changing its rates of binding to and unbinding from the MT.
The interplay between motor traffic and depolymerization kinetics at the mi-
crotubule plus-end leads to strong correlations between the depolymerization dy-
namics and density profiles of depolymerases. The plus-end acts as a bottleneck
and crowding effects cause traffic jams. We find two qualitatively distinct regimes:
18
Motor densities below a critical threshold value, ρ∗
La = δτ, always show a local
spike-like perturbation at the plus-end, the extent of which is the size of a het-
erodimer. Above this threshold density, macroscopic traffic jams may emerge.
These distinct density profiles at the plus-end affect the depolymerization speed
and the end-residence time in qualitatively different ways. A quantitative analy-
sis of the model using stochastic simulations as well as analytical calculations led
to the following main results: The end-residence time of a depolymerase strongly
depends on the degree of cooperativity. Whereas for noncooperative depolymeriza-
tion kinetics the end-residence time is given by the microscopic depolymerization
rate δ, it is density-dependent in the fully cooperative case: Increasing the Lang-
muir density above the threshold value ρ∗
La = δτ, the end-residence time changes
from being inversely proportional to the density ρLa to a constant value δ−1. These
results suggest an interesting way to determine the cooperativity of depolymeriza-
tion kinetics and measure the value of the depolymerization rate δ. Although when
the concentration c is increased, the end-residence time should be independent of
concentration for noncooperative kinetics, it should strongly depend on concen-
tration in the cooperative case. Experimental evidence points toward the latter
(23).
In contrast, the depolymerization speed does not depend on the degree of coop-
erativity of the depolymerization kinetics. Noncooperative and fully cooperative
versions of the model give identical results. As a function of depolymerase concen-
tration and the MT length, the depolymerization dynamics exhibits two qualita-
tively distinct regimes: The depolymerization speed is either density-limited and
determined by the bulk density of molecular motors, ρ−(x), or rate-limited and
dictated by the value of the microscopic depolymerization rate, δ. Both regimes
emerge due to crowding of molecular motors at the plus-end which acts as a bot-
tleneck for molecular traffic.
Density-limited regimes are correlated with microscopic traffic jams ("spikes")
at the plus-end: The density profile self-organizes into a shape that cancels out all
the effects of the depolymerization kinetics such that the depolymerization speed is
solely determined by the bulk motor density, ρ−(x), and the average motor speed,
v. Note that only in this regime length-dependent regulation is possible since the
density changes over the MT length. As emphasized above, if the depolymerization
rate δ is larger than the hopping rate of the molecular motors, δ > ν, this remains
the only regime of depolymerization dynamics. Then, the depolymerization speed
is limited by the velocity of the plus-end directed motors, which is in accordance
with recent experimental findings for Kip3p (23). In a parameter regime where
motors depolymerize more slowly than they walk, δ < ν, there is a second rate-
limited regime above the threshold density ρ∗
La and for microtubules longer than
some threshold length (cid:96)reg where Vdepol = aδ.
In this regime the plus-end acts
19
as a strong bottleneck for molecular traffic. This causes a macroscopic traffic jam
such that the motor density steeply rises to full occupation of all lattice sites at the
plus-end of the microtubule. The cellular system sacrifices its capability to regulate
the speed of depolymerization and only regains it once the MT length falls below
(cid:96)reg, where the depolymerization speed again becomes density-regulated. From an
evolutionary perspective one might speculate that the system has evolved towards
δ = ν, because this would allow regulation of the depolymerization dynamics over
the broadest possible range.
Beyond these observations, other predictions of our stochastic model can be
put to test in experiments. By varying the motor concentration, two interest-
ing observations could be made: First the phase diagram for the density profiles
at the minus-end could be scrutinized experimentally. Second, the predictions
on the density-profiles at the plus-end and their predicted strong correlations to
the macroscopic depolymerization dynamics might be accessible to single-molecule
studies. Manipulation of the molecular properties of the motor (e.g., the run
length, attachment rate (42), average speed and depolymerization rate) would
change the intrinsic biochemical rates of the system and potentially lead to new
parameter regimes. In addition, our results regarding length- and concentration
dependence of the depolymerization process might be relevant in vivo, e.g., for
mitotic chromosome alignment (18).
In our theoretical studies we explored the
full parameter range, and therefore clear predictions are available for comparison.
We believe that in a more general context, our theoretical work provides new
conceptual insights into the role of collective and cooperative effects in micro-
tubule assembly and disassembly dynamics. Future research could focus on the
antagonism between polymerases and depolymerases (12, 43, 44), spontaneous MT
dynamics mediated by GTP-hydrolysis, the abundance of molecular motors in a
cell, or more-detailed modeling of molecular motors (26). This may finally lead to
a molecular understanding of the regulatory mechanisms of cellular processes in
which MT dynamics plays a central role.
ACKNOWLEDGMENTS
The authors thank Cécile Leduc for discussions, the authors of (23) for kindly
providing their data, Ulrich Gerland, Günther Woehlke and Jonas Cremer for
critical reading of the original manuscript, Anton Winkler for helpful suggestions
on the revised manuscript and Andrej Vilfan for drawing Fig. 1A. This project was
supported by the Deutsche Forschungsgemeinschaft in the framework of the SFB
863 and the German Excellence Initiative via the program "Nanosystems Initiative
Munich" (NIM).
20
References
[1] Hayles, J., and P. Nurse. 2001. A journey into space. Nat. Rev. Mol. Cell
Biol. 2:647 -- 656.
[2] Tolić-Nørrelykke, I. M. 2010. Force and length regulation in the microtubule
cytoskeleton: lessons from fission yeast. Curr. Opin. Cell Biol. 22:21 -- 28.
[3] Sharp, D., G. Rogers, and J. Scholey. 2000. Microtubule motors in mitosis.
Nature. 407:41 -- 47.
[4] Karsenti, E., and I. Vernos. 2001. Cell cycle - the mitotic spindle: A self-made
machine. Science. 294:543 -- 547.
[5] Eggert, U. S., T. J. Mitchison, and C. M. Field. 2006. Animal cytokinesis:
From parts list to mechanisms. Annu. Rev. Biochem. 75:543 -- 566.
[6] Hirokawa, N., Y. Noda, Y. Tanaka, and S. Niwa. 2009. Cytoskeletal motors:
Kinesin superfamily motor proteins and intracellular transport. Nat. Rev.
Mol. Cell Biol. 10:682 -- 696.
[7] Mitchison, T., and M. Kirschner. 1984. Dynamic instability of microtubule
growth. Nature. 312:237 -- 242.
[8] Dogterom, M., and S. Leibler. 1993. Physical aspects of the growth and
regulation of microtubule structures. Phys. Rev. Lett. 70:1347 -- 1350.
[9] Desai, A., and T. Mitchison. 1997. Microtubule polymerization dynamics.
Annu. Rev. Cell Dev. Biol. 13:83 -- 117.
[10] Howard, J., and A. Hyman. 2003. Dynamics and mechanics of the microtubule
plus end. Nature. 422:753 -- 758.
[11] Wordeman, L. 2005. Microtubule-depolymerizing kinesins. Curr. Opin. Cell
Biol. 17:82 -- 88.
[12] Howard, J., and A. A. Hyman. 2007. Microtubule polymerases and depoly-
merases. Curr. Opin. Cell Biol. 19:31 -- 35.
[13] Howard, J., and A. A. Hyman. 2009. Growth, fluctuation and switching at
microtubule plus ends. Nat. Rev. Mol. Cell Biol. 10:569 -- 574.
[14] Helenius, J., G. Brouhard, Y. Kalaidzidis, S. Diez, and J. Howard. 2006.
The depolymerizing kinesin MCAK uses lattice diffusion to rapidly target
microtubule ends. Nature. 441:115 -- 119.
21
[15] Varga, V., J. Helenius, K. Tanaka, A. A. Hyman, T. U. Tanaka, and
J. Howard. 2006. Yeast kinesin-8 depolymerizes microtubules in a length-
dependent manner. Nat. Cell Biol. 8:957 -- 962.
[16] Gupta, M. L., P. Carvalho, D. M. Roof, and D. Pellman. 2006. Plus end-
specific depolymerase activity of Kip3, a kinesin-8 protein, explains its role in
positioning the yeast mitotic spindle. Nat. Cell Biol. 8:913 -- 923.
[17] Mayr, M. I., S. Hümmer, J. Bormann, T. Grüner, S. Adio, G. Woehlke,
and T. U. Mayer. 2007. The human kinesin Kif18a is a motile microtubule
depolymerase essential for chromosome congression. Curr. Biol. 17:488 -- 498.
[18] Stumpff, J., G. V. Dassow, M. Wagenbach, C. Asbury, and L. Wordeman.
2008. The kinesin-8 motor Kif18a suppresses kinetochore movements to con-
trol mitotic chromosome alignment. Dev. Cell. 14:252 -- 262.
[19] Du, Y., C. A. English, and R. Ohi. 2010. The kinesin-8 Kif18a dampens
microtubule plus-end dynamics. Curr. Biol. 20:374 -- 380.
[20] Unsworth, A., H. Masuda, S. Dhut, and T. Toda. 2008. Fission yeast kinesin-8
Klp5 and Klp6 are interdependent for mitotic nuclear retention and required
for proper microtubule dynamics. Mol. Biol. Cell. 19:5104 -- 5115.
[21] Tischer, C., D. Brunner, and M. Dogterom. 2009. Force-and kinesin-8-
dependent effects in the spatial regulation of fission yeast microtubule dy-
namics. Mol. Syst. Biol. 5:250.
[22] Grissom, P., T. Fiedler, E. Grishchuk, D. Nicastro, R. West, and J. R. McIn-
tosh. 2009. Kinesin-8 from fission yeast: A heterodimeric, plus-end-directed
motor that can couple microtubule depolymerization to cargo movement. Mol.
Biol. Cell. 20:963.
[23] Varga, V., C. Leduc, V. Bormuth, S. Diez, and J. Howard. 2009. Kinesin-8
motors act cooperatively to mediate length-dependent microtubule depoly-
merization. Cell. 138:1174 -- 1183.
[24] Gardner, M. K., D. C. Bouck, L. V. Paliulis, J. B. Meehl, E. T. O'Toole,
J. Haase, A. Soubry, A. P. Joglekar, M. Winey, E. D. Salmon, K. Bloom,
and D. J. Odde. 2008. Chromosome congression by kinesin-5 motor-mediated
disassembly of longer kinetochore microtubules. Cell. 135:894 -- 906.
[25] Foethke, D., T. Makushok, D. Brunner, and F. Nédélec. 2009. Force- and
length-dependent catastrophe activities explain interphase microtubule orga-
nization in fission yeast. Mol. Syst. Biol. 5:241.
22
[26] Klumpp, S., Y. Chai, and R. Lipowsky. 2008. Effects of the chemomechanical
stepping cycle on the traffic of molecular motors. Phys. Rev. E. 78:041909.
[27] Howard, J. 1996. The movement of kinesin along microtubules. Annu. Rev.
Physiol. 58:703 -- 729.
[28] Ray, S., E. Meyhöfer, R. Milligan, and J. Howard. 1993. Kinesin follows the
microtubule protofilament axis. J. Cell Biol. 121:1083 -- 1093.
[29] Parmeggiani, A., T. Franosch, and E. Frey. 2003. Phase coexistence in driven
one-dimensional transport. Phys. Rev. Lett. 90:086601.
[30] Parmeggiani, A., T. Franosch, and E. Frey. 2004. Totally asymmetric simple
exclusion process with langmuir kinetics. Phys. Rev. E. 70:046101.
[31] Lipowsky, R., S. Klumpp, and T. Nieuwenhuizen. 2001. Random walks
of cytoskeletal motors in open and closed compartments. Phys. Rev. Lett.
87:108101.
[32] Klumpp, S., and R. Lipowsky. 2003. Traffic of molecular motors through
tube-like compartments. J. Stat. Phys. 113:233 -- 268.
[33] Pierobon, P., M. Mobilia, R. Kouyos, and E. Frey. 2006. Bottleneck-induced
transitions in a minimal model for intracellular transport. Phys. Rev. E.
74:031906.
[34] Telley, I. A., P. Bieling, and T. Surrey. 2009. Obstacles on the microtubule
reduce the processivity of Kinesin-1 in a minimal in vitro system and in cell
extract. Biophys. J. 96:3341 -- 3353.
[35] Govindan, B. S., M. Gopalakrishnan, and D. Chowdhury. 2008. Length control
of microtubules by depolymerizing motor proteins. Europhys. Lett. 83:40006.
[36] Brun, L., B. Rupp, J. J. Ward, and F. Nedelec. 2009. A theory of microtubule
catastrophes and their regulation. Proc. Natl. Acad. Sci. USA. 106:21173 --
21178.
[37] Hough, L. E., A. Schwabe, M. A. Glaser, J. R. McIntosh, and M. D. Better-
ton. 2009. Microtubule depolymerization by the kinesin-8 motor Kip3p: A
mathematical model. Biophys. J. 96:3050 -- 3064.
[38] Klein, G. A., K. Kruse, G. Cuniberti, and F. Jülicher. 2005. Filament depoly-
merization by motor molecules. Phys. Rev. Lett. 94:108102.
23
[39] Vilfan, A., E. Frey, F. Schwabl, M. Thormahlen, Y. Song, and E. Mandelkow.
2001. Dynamics and cooperativity of microtubule decoration by the motor
protein kinesin. J. Mol. Biol. 312:1011 -- 1026.
[40] Frey, E., and A. Vilfan. 2002. Anomalous relaxation kinetics of biological
lattice-ligand binding models. Chem. Phys. 284:287 -- 310.
[41] Frey, E., A. Parmeggiani, and T. Franosch. 2004. Collective phenomena in
intracellular processes. Genome Informatics. 15(1):46 -- 55.
[42] Cooper, J. R., M. Wagenbach, C. L. Asbury, and L. Wordeman. 2010. Catal-
ysis of the microtubule on-rate is the major parameter regulating the depoly-
merase activity of MCAK. Nat. Struct. Mol. Biol. 17:77 -- 82.
[43] Kinoshita, K., I. Arnal, A. Desai, D. Drechsel, and A. Hyman. 2001. Recon-
stitution of physiological microtubule dynamics using purified components.
Science. 294:1340 -- 1343.
[44] Brouhard, G. J., J. H. Stear, T. L. Noetzel, J. Al-Bassam, K. Kinoshita, S. C.
Harrison, J. Howard, and A. A. Hyman. 2008. XMAP215 is a processive
microtubule polymerase. Cell. 132:79 -- 88.
24
Supporting Material to "Crowding of molecular motors
determines microtubule depolymerization"
Louis Reese, Anna Melbinger and Erwin Frey
(Corresponding author. Email: [email protected])
Arnold Sommerfeld Center for Theoretical Physics and Center for NanoScience,
Department of Physics, Ludwig-Maximilians-Universität München,
Theresienstrasse 37, 80333 Munich, Germany
In this Supporting Material, details concerning the mathematical formulation
and the stochastic simulations are given. In particular, the density profiles and
the domain wall positions at the minus- and the plus-end are derived analytically.
Further, some additional results are provided: (i) We show that the shapes of MT
depolymerization curves (kymographs) are to a large extent independent of the
choice of the depolymerization rate δ; see Fig. S1. (ii) Analytical and numerical
results from our theory are compared to experimental data on the relation between
depolymerization speed and motor current (1); see Fig. S2.
Mathematical formulation
In this article, we employ a lattice gas model. Its state is described by a set of
occupation numbers ni ∈ {0, 1} where i = 1, . . . , L denotes the lattice sites. In
contrast to the notation in the main text, we here choose units of length and time
such that the hopping rate from site to site and the lattice constant are both set
to one. For an analytical description of the steady state density profiles of the
molecular motors along the MT we consider the ensemble-averaged densities and
currents:
ρi := (cid:104)ni(cid:105) ,
Ji := (cid:104)ni(1 − ni+1)(cid:105) .
(S1)
(S2)
Note that the current Ji accounts for particle exclusion: a particle at site i moves
to site i + 1 at rate ν = 1 only if site i + 1 is unoccupied. The steady state results
from a local balance between the transport current (2),
:= (cid:104)ni−1(1 − ni)(cid:105) − (cid:104)ni(1 − ni+1)(cid:105) ,
J T
i
the particle exchange with the bulk,
:= c ωa(cid:104)1 − ni(cid:105) − ωd(cid:104)ni(cid:105) ,
J La
i
25
(S3)
(S4)
and the depolymerization current, which sets the boundary condition at the plus-
end. We now perform a mean-field approximation, where all spatial correlations
are neglected, and a continuum limit keeping only the leading order terms (3).
Then, the transport current simplifies to,
J T (x) =(cid:0)2ρ(x) − 1(cid:1)∂xρ(x) ,
(S5)
(S6)
i.e. the transport current is proportional to the density gradient like a diffusion
current in Fick's law but modified with a density-dependent prefactor which reflects
site-exclusion between motors. The Langmuir current is given by
(cid:0)1 − ρ(x)(cid:1) − ωdρ(x) .
J La(x) = c ωa
Density profiles at the minus-end
Within the above introduced framework the motor density profiles on the MT
can be calculated analytically.
In particular, the domain wall position, can be
derived exactly as well as upon employing a linear approximation for the density
profile close to the minus-end. For simplicity, we first consider the latter, especially
because its results approximate the exact solution rather well over a broad range
of parameters.
Linear approximation
In the immediate vicinity of the minus-end (x = 0) the density is small such that
the full equation for the current balance, J T + J La = 0,
(2ρ(x) − 1)∂xρ(x) + c ωa(1 − ρ(x)) − ωdρ(x) = 0 ,
(S7)
reduces to ∂xρ = c ωa, which is solved by a linear (antenna) profile:
ρ−(x) ≈ c ωax .
(S8)
At sufficiently large distances from the minus-end the density profile becomes flat.
Therefore, J T vanishes and the system is dominated by the Langmuir kinetics,
J La = 0. Then, an asymptotic solution of Eq. S7 is given by the Langmuir density
ρLa =
K
1 + K
=
c ωa
c ωa + ωd
.
(S9)
The full density profile is obtained by concatenating the antenna profile and the
flat Langmuir profile such that the (local) current is continuous along the MT.
There are two qualitatively distinct scenarios. For low bulk concentrations of
molecular motors, c, the antenna profile matches the asymptotic Langmuir density
26
continuously resulting in a wedge-like profile; compare Fig. 2E in the main text.
Approximately, the matching point, ρ−(d−
w) = ρLa, is
w ≈
(cid:96)−
K
c ωa(K + 1)
.
(S10)
In contrast, above a certain threshold value for the bulk concentration, determined
by K−
c = 1, the two profiles can no longer be matched continuously and the
density profile displays a localized discontinuity (2), also termed a "domain wall"
(DW). Its position is determined by a local current continuity condition (2, 3),
ρ−(d−) = 1 − ρLa, and can again be estimated using the linear antenna profile:
Taken together, Eq. (5) from the main text is obtained,
(cid:96)− ≈
1
c ωa(K + 1)
.
(cid:40) 1
(cid:96)− =
ωd(K+1)
1
c ωa(K+1)
for K < 1 ,
for K > 1 .
(S11)
(S12)
Exact solution and domain wall position
To obtain the full solution Eq. S7 has to be solved as already demonstrated in
K−1 (2ρ(x) − 1)−
Ref. (3). Introducing a rescaled density at the minus-end σ−(x) = K+1
1 in Eq. S7 a transformed differential equation can be obtained
∂xσ−(x) + ∂x lnσ(x)− = ωd
(K + 1)2
K − 1
,
which is mathematically equivalent to Eq. S7 and can be solved analytically
σ−(x) = W−1 (−Y−(x)) .
(S13)
Here W−1 is the second real branch of the Lambert W -function (4) and Y−(x)
reads (3)
.
(K + 1)2
K − 1
x − 2K
K − 1
ωd
Y−(x) =
(S14)
Herein the boundary condition ρ−(0) = 0 corresponding to σ−(0) = −2K/(K − 1)
has already been accounted for. The local current condition for the domain wall,
ρ−(d−) = 1− ρLa which corresponds to σ−(d−) = −2 for the rescaled density, now
enables us to calculate the DW position. Combining this condition with Eqs. S13
and S14 leads to
d−(ωd, K) =
2 + (K − 1) ln(1 − 1/K)
(S15)
.
ωd(K + 1)2
(cid:12)(cid:12)(cid:12)(cid:12) −2K
K − 1
(cid:12)(cid:12)(cid:12)(cid:12) exp
(cid:26)
(cid:27)
27
Density profiles at the plus-end
Analogously to the minus-end, we now evaluate the density profiles and ensuing
DW at the plus-end. Because the tip steadily depolymerizes, the calculations have
to be performed in a comoving frame which is introduced first.
Comoving frame
In the comoving frame, the above defined last lattice site L, i.e. the plus-end,
is defined as the first site of the MT. This is equivalent to reverting the motor
movement. Since molecular motors in the comoving frame move towards the first
site of the lattice, the transport current changes sign J T
i = −J T ;Co
i
J T ;Co
i
:= (cid:104)ni+1(1 − ni)(cid:105) − (cid:104)ni(1 − ni−1)(cid:105) ,
using the mean-field approximation as introduced above this leads to
J T ;Co(x) = −(2ρ − 1)∂xρ(x) .
(S16)
(S17)
The particle adsorption/desorption current J La
is unaffected. However, there is
another contribution to the current balance in the comoving frame due to depoly-
merization: Similar to the above definitions of the local currents, a local current
which accounts for depolymerization in the comoving frame arises
i
i = δ(ni+1 − ni) .
J Co
Employing a mean-field approximation this expression simplifies to,
J Co(x) = δ∂xρ(x) .
(S18)
(S19)
This current term can be understood as follows. Due to the depolymerizing activity
of a motor at the plus-end, in the comoving frame all motors on the MT simul-
taneously approach the plus-end. In summary, by introducing a comoving frame
the mean-field equation for the density at the MT plus-end ρ+(x) is obtained. In
the steady state it reads
(2ρ+ − 1 − δ)∂xρ+ + cωa(1 − ρ+) − ωdρ+ = 0 .
(S20)
Density profiles
The above equation is solved in close analogy to Eq. S7. In terms of a rescaled
density
σ+(x) =
2ρ+(x) − 2 K
K+1 − (1 + δ)
2 K
K+1
,
28
(S21)
the rescaled differential equation reads
+(x) + ∂x lnσ+(x) =
σ(cid:48)
ωd(K + 1)2
K − 1 − (K + 1)δ
.
(S22)
The exact solutions to this equation are compared to stochastic simulations in the
main text (Fig. 5A).
The solutions of this equation for δ = 0 are discussed in (3) and in parts above.
However, in the case of depolymerization, i.e.
for δ > 0, two special solutions
exist: Depending on the density of motors on the MT two classes of solutions for
the density at the plus-end ρ+(x) can be distinguished. These are wedge-like or
traffic jam density profiles; see Fig. 5A in the main text. The boundary condition
for these qualitatively distinct density profiles is ρ+(L) = 1. Defining
Y (x) = σ+(L) exp
ωd(K + 1)2
K − 1 − δ(K + 1)
(x − L) + σ+(L)
,
(S23)
(cid:27)
(cid:26)
(cid:40)
the two solutions for density profiles in the main text (black lines in Fig. 5A) are
ρ(x) =
ρLa + 1
ρLa + 1
2(2ρLa − (1 + δ))W0(Y (x))
2(2ρLa − (1 + δ))W−1(−Y (x))
wedge-like
traffic jam.
(S24)
W0 and W−1 denote the first and the second real branch of the Lambert function.
The reason for the form of these two solutions is the bottleneck (5) arising due to
depolymerization (see main text). This bottleneck fixes the value of the tip density
to its maximum ρ+(L) = 1 for ρLa > δ. The transition from a traffic jam to a
wedge-like density profile is thus not boundary-induced, i.e. due to a particular
value of ρ+(L) < 1, but may be attributed to the depolymerizing activity of motors
at the plus-end. As discussed in the main text, this transition is sharp and can be
quantified in terms of ρLa and δ; see Fig. 5B.
Domain wall position at the plus-end
As already shown in the main text, microscopic jams can substantially influence
the depolymerization dynamics. For large bulk concentrations, this perturbation
no longer remains a local spike, but affects the profile on a macroscopic scale
(5). Because the perturbation is macroscopic we can again use a hydrodynamic
description, now with the boundary condition ρ+(L) = 1. Close to the plus-end
Eq. S20 gives an approximately linear profile
ρ+(x) ≈ 1 − ωd
1 − δ
(L − x) .
(S25)
The slope increases with increasing depolymerization rate δ, c.f. Fig. 5A in the
main text. In close analogy with the discussion for the minus-end there are two
29
scenarios for concatenating this linear profile with the Langmuir density. Here, for
large enough Langmuir density and/or small enough depolymerization rates, we
obtain a wedge-like profile with a matching point given by ρ+(d+
w) = ρLa:
;
(S26)
w ≈ L − 1 − δ
d+
ωd(1 + K)
(cid:20) K
1 + K
(cid:21) 1 − δ
ωd
− δ
d+ ≈ L −
compare the green curves in Fig. 5A in the main text. Upon increasing the depoly-
merization rate or decreasing the Langmuir density a DW emerges whose position
can be determined using current conservation, ρ+(d+) = 1 − ρLa + δ:
;
(S27)
compare the red curves in Fig. 5A. The DW is most pronounced for K < 1. The
height of the DW vanishes as K approaches the threshold value
K +
c =
1 + δ
1 − δ
(S28)
from below. Equivalently, for a given K, the critical depolymerization rate reads
K − 1
K + 1
δc =
= 2ρLa − 1 =
c ωa − ωd
c ωa + ωd
.
(S29)
Together with the condition for spikes, ρLa < δ, this relation organizes the shapes
of the density profiles into three classes: Microscopic jams at the tip, wedge profiles
and DW profiles.
Depolymerization dynamics of the antenna profile
In the main text we have discussed how a spatially uniform density ρLa affects
depolymerization dynamics. Here we briefly show that our approach is also ap-
plicable to linear antenna profiles, i.e.
for MTs shorter than a certain threshold
length, L < (cid:96)−, cf. Eq. (S12) and main text. Just as in the main text we equate
the particle loss current due to depolymerization,
Jdepol(x) = [1 − ρ−(x)] ρ+(x)δ ,
and the particle flux towards the plus-end,
J−(x) = [1 − ρ−(x)] ρ−(x) ,
and find
(cid:40) ρ−(x)
ρ+(x) =
δ
1
for L < δ
cωa
for L > δ
cωa
,
.
30
(S30)
(S31)
(S32)
Since, according to Eq. (2) in the main text, the density at the plus-end deter-
mines the depolymerization speed, Vdepol = δρ+(x), the position-dependence of the
tip density maps to a length-dependence of the polymerization speed. For MTs
shorter than (cid:96)− but longer than a certain depolymerization length (cid:96)d = δ/cωa the
depolymerization speed is length-independent
Vdepol = ρ+δ = δ .
(S33)
Analogously to the result for constant bulk densities, this result shows that for MTs
longer than the depolymerization length (cid:96)d the dynamics of depolymerization of
the antenna profile can not be distinguished from the dynamics as induced by a flat
density profile. In contrast, at a MT length shorter than (cid:96)d the depolymerization
speed becomes length-dependent and follows the shape of the antenna density
profile ρ−(x):
Vdepol(L) = ρ+(L)δ = ρ−(L) .
(S34)
These results generalize the rate-limited and density-limited regimes discussed in
the main text to non-uniform densities. Moreover, they show that once filaments
become shorter than (cid:96)−, i.e. the density profile is antenna-like, there is a second
spike-induced length scale (cid:96)d which is the relevant length scale for the onset of
length-dependent depolymerization of MTs.
Combining these results with the analogous conditions for the Langmuir plateau
discussed in the main text, leads to the depolymerization regimes summarized in
Fig. 5C and Table 1. Simply put, the depolymerization dynamics changes from
rate-limited to density-limited when the bulk density falls below the threshold den-
sity δ: ρ−(L) ≤ δ. For Langmuir densities below the threshold density, ρLa < δ,
the bulk density remains below the threshold density for the whole MT length
such that the depolymerization dynamics is always density-limited and given by:
Vdepol = ρ−(L). This corresponds to regime (α) in Fig. 5C. For Langmuir densities
above the threshold, ρLa > δ, the depolymerization dynamics is rate-limited in
the Langmuir plateau and given by Vdepol = δ. In the antenna-like regime of the
density profile, i.e. for L ≤ (cid:96)−, we have to distinguish between two cases: (i) K > 1
(ρLa > 0.5) where the bulk density profile exhibits a domain wall, and (ii) K < 1
(ρLa < 0.5) where the bulk density profile is wedge-like. In the latter case, the
bulk-density profile changes slowly and hence remains above the threshold δ for
some time even below L = (cid:96)−. Only for MTs shorter than (cid:96)d, given by ρ−((cid:96)d) = δ,
the dynamics changes from rate- to density-limited. This corresponds to regime
(γ) in Fig. 5C. In contrast, for K > 1 (ρLa > 0.5), the bulk density ρ−(x) ex-
hibits a discontinuous jump from the Langmuir density to 1 − ρLa right at (cid:96)−. If
this maximum value of the antenna-like profile is less than the threshold density,
1 − ρLa < δ, then the depolymerization dynamics discontinuously switches from
31
rate-limited to density-limited. This defines regime (β) in Fig. 5C. Otherwise, if
1 − ρLa > δ, we are back to regime (γ). In summary, all regimes show a constant
polymerization speed for long MTs in the Langmuir plateau. Depending on the
relative magnitude of the Langmuir density and the depolymerization rate this
regime is either density-limited and given by ρLa or rate-limited and given by δ, cf.
second column in Table 1. In all scenarios the dynamics becomes length-dependent
at some scale which is, however, different. While for regimes α and β, it coincides
with the beginning of the antenna-like density profile (cid:96)−, it is given by (cid:96)d for regime
(γ) cf. third column in Table 1.
Regime
(α)
(β)
(γ)
Condition
ρLa < δ
ρLa > δ > 1 − ρLa
else
Constant Vdepol Critical MT length
Vdepol = ρLa
Vdepol = δ
Vdepol = δ
(cid:96)−
(cid:96)−
(cid:96)d
Table 1: Summary of the similarities and differences in the depolymerization
regimes (α), (β) and (γ). While for MTs longer than a critical length (third col-
umn) the depolymerization speed is constant (second column), it becomes length-
dependent below this critical length and is then given by Vdepol = ρ−(L).
Numerical implementation
The stochastic dynamics of the individual-based model was simulated using a
Gillespie algorithm (7) and employing the rates introduced above. Note that this
method provides the mathematically exact stochastic dynamics. This is essential
for the investigation of dynamic phenomena like length-dependent shortening.
In Fig. 2A-D., our simulations of the motor traffic started from an initial con-
dition where the MT lattice was empty and subsequently filled up with motors
triggering the depolymerization dynamics. To visualize time-dependent MT length
and motor densities in one kymograph we implemented 14 protofilaments and av-
eraged the motor intensities and MT lengths, see Fig. 2A-C. In detail, the visual-
ization of kymographs was achieved as described in the following. From stochastic
simulation data of the MT, each second the occupation numbers of motors along
, where j indexes the 14 protofilaments of the MT, were evaluated and
the MT n(j)
(cid:80)14
(cid:80)14
converted to RGB color values:
j=1 1−n(j)
(S35)
These values display the density of motors as green and the uncovered MT surface
as red. Hence, if the MT is completely empty it is red, while at complete motor
coverage it is green.
, Bi = 0.
j=1 n(j)
14
i
Ri =
i
, Gi =
14
i
32
Steady state motor densities as shown in Figs. 2E and 5A were obtained by
time-averaging over 104 independent realizations after an equilibration time of
2000 time steps; note that for a constant lattice size time and ensemble averages
yield identical results (2). In Fig. 5A, the density profiles were recorded in the
comoving frame of the MT plus-end, while density profiles in Fig. 2E were recorded
in the rest frame of the MT minus-end. In Fig. 2E, the density profiles resulting
from the minus end without any influences from the plus end are shown. This
can be viewed as an infinitely long lattice. To simulate such a lattice, we chose
the following boundary conditions. We neglected depolymerization as it arises
at infinity. Further, we set the exiting rate for motors at the last site equal to
1 − ρLa. Then, the transport behavior at the tip is the same as on the lattice, if
the Langmuir density is reached, ρ(x) = ρLa.
In the second part of the article we focus on the dependence of the depolymer-
ization speed on the motor density. To this end, simulations were performed in
a comoving frame where the density at the minus-end was fixed to the Langmuir
density. This was achieved by extending the lattice one site to left with each de-
polymerization step and filling the thereby created site with the probability ρLa.
This procedure may also be interpreted as an infinite MT allowing to observe motor
dynamics at the MT tip without perturbations arising from the length-dependent
depolymerization regime.
We measured the mean end-residence time of individual motors at the plus-end
τres and the mean lifetime of the terminal tubulin dimer τd. Data of these were
obtained by averaging over 104 time steps τ, cf. Figs. 4 and 6.
How kymographs become independent of the de-
polymerization rate
In Fig. S1 we provide data that explicitly shows the parameter independence of
MT depolymerization. This results has been generalized in the main text to all
possible motor concentrations and depolymerization rates.
Comparison with experiments: dependence of the
depolymerization speed on the bulk flux and bulk
density
Experimentally it was found that the depolymerization speed is linearly correlated
with the flux of molecular motors towards the plus-end (1). We have collected
data from our simulations similar to experiments. Figure S2 shows a scatter plot
33
FIGURE S1: Kymographs as they become independent of the depolymerization
rate δ. Time-space plots of depolymerizing MTs for different depolymerization
rates are shown, ranging from δ = 0.1 ν (red) to 1.0 ν (blue), the latter value cor-
responds to the motor speed of v = 6.35 sites s−1. For slow depolymerization rate,
δ (cid:47) 0.5, the depolymerization speed is related to the microscopic depolymeriza-
tion rate δ, whereas for rates δ (cid:39) 0.5 the depolymerization speed is independent
of the depolymerization rate but depends on the density of motors on the MT, as
outlined in the main text.
for the depolymerization speed as a function of the bulk flux of motors, JLa, for
two values of the microscopic depolymerization rate δ.
The noise in the ensemble of realizations has two sources. The bulk current
fluctuates since the Langmuir kinetics responsible for the bulk density ρLa is a
stochastic process. The depolymerization speeds vary from realization to realiza-
tion because the depolymerization kinetics is a Poisson-like process. Also shown in
Fig. S2 are ensemble averages. These mean values, as predicted in the main text,
show the following behavior. For a macroscopic depolymerization speed lower than
the depolymerization rate, ∆L/∆T < δa, it is density-limited and identical to the
bulk density:
∆L
∆T
= ρLav
for ρLa < δτ .
(S36)
Rewriting this relation in terms of the bulk current means that the data should
fall on the parabola displayed as the solid curve in Fig. S2. For low densities,
ρLa (cid:47) 0.25, where crowding effects are weak, this implies ∆L/∆T = ρLav ≈ JLa as
observed experimentally (1); see Fig. S2.
As the bulk density is increased two things happen. First, crowding effects
become important invalidating the linear relationship between bulk current and
34
05101520253035Time t [min]020040060080010001200MT length L [sites]δ0=1.0δ0=0.9δ0=0.8δ0=0.7δ0=0.6δ0=0.5δ0=0.4δ0=0.3δ0=0.2δ0=0.1FIGURE S2: Depolymerization speed ∆L/∆T as a function of the bulk motor cur-
rent on the MT, JLa. Data from stochastic simulations for two different depoly-
merization rates δ = 0.5 ν (◦) and 0.8 ν ((cid:47)) are shown, each for a set of Langmuir
densities (colors) corresponding to concentrations c = 0.16 . . . 13 nM. To illustrate
the effect of statistics, small symbols show individual measurements as obtained
from 103 fixed time measurements ∆T = 500τ, while large symbols indicate their
mean (cid:104)∆L/∆T(cid:105). Good agreement with experimental data ((cid:50)) measured in the low
density regime is found (1). Here, the theoretical prediction given by JLa (solid)
is hardly discernible from a linear best fit to experiments (dashed).
depolymerization speed. It would be interesting to test our prediction that the
depolymerization speed is linear in the bulk density by using higher motor con-
centrations or changed biochemical rates such that K becomes significantly larger
than 1.
Second, if ρLa > δτ, the depolymerization speed becomes rate-limited:
∆L
∆T
= aδ
for ρLa > δτ .
(S37)
This puts an obvious upper bound on the depolymerization speed.
It cannot
become larger than the microscopic rate of depolymerization at the plus-end. If
the depolymerization rate is larger than the hopping rate of the molecular motors,
δ > ν, the depolymerization speed is, for all possible values of the bulk density,
strictly given by the bulk density.
References
[1] Varga, V., C. Leduc, V. Bormuth, S. Diez, and J. Howard. 2009. Kinesin-8
motors act cooperatively to mediate length-dependent microtubule depolymer-
35
ization. Cell. 138:1174 -- 1183.
[2] Parmeggiani, A., T. Franosch, and E. Frey. 2003. Phase coexistence in driven
one-dimensional transport. Phys. Rev. Lett. 90:086601.
[3] Parmeggiani, A., T. Franosch, and E. Frey. 2004. Totally asymmetric simple
exclusion process with langmuir kinetics. Phys. Rev. E. 70:046101.
[4] Corless, R., G. Gonnet, D. Hare, D. Jeffrey, and D. Knuth. 1996. On the
lambert w function. Adv. Comput. Math. 5:329 -- 359.
[5] Pierobon, P., M. Mobilia, R. Kouyos, and E. Frey. 2006. Bottleneck-induced
transitions in a minimal model for intracellular transport. Phys. Rev. E.
74:031906.
[6] Varga, V., J. Helenius, K. Tanaka, A. A. Hyman, T. U. Tanaka, and J. Howard.
2006. Yeast kinesin-8 depolymerizes microtubules in a length-dependent man-
ner. Nat. Cell Biol. 8:957 -- 962.
[7] Gillespie, D. T. 1976. Stochastic simulations of chemical processes. J. Comp.
Phys. 22:403 -- 434.
36
|
1607.06640 | 1 | 1607 | 2016-07-22T11:51:24 | Ephemeral protein binding to DNA shapes stable nuclear bodies and chromatin domains | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech",
"q-bio.SC"
] | Fluorescence microscopy reveals that the contents of many (membrane-free) nuclear "bodies" exchange rapidly with the soluble pool whilst the underlying structure persists; such observations await a satisfactory biophysical explanation. To shed light on this, we perform large-scale Brownian dynamics simulations of a chromatin fiber interacting with an ensemble of (multivalent) DNA-binding proteins; these proteins switch between two states -- active (binding) and inactive (non-binding). This system provides a model for any DNA-binding protein that can be modified post-translationally to change its affinity for DNA (e.g., like the phosphorylation of a transcription factor). Due to this out-of-equilibrium process, proteins spontaneously assemble into clusters of self-limiting size, as individual proteins in a cluster exchange with the soluble pool with kinetics like those seen in photo-bleaching experiments. This behavior contrasts sharply with that exhibited by "equilibrium", or non-switching, proteins that exist only in the binding state; when these bind to DNA non-specifically, they form clusters that grow indefinitely in size. Our results point to post-translational modification of chromatin-bridging proteins as a generic mechanism driving the self-assembly of highly dynamic, non-equilibrium, protein clusters with the properties of nuclear bodies. Such active modification also reshapes intra-chromatin contacts to give networks resembling those seen in topologically-associating domains, as switching markedly favors local (short-range) contacts over distant ones. | physics.bio-ph | physics |
Ephemeral protein binding to DNA shapes stable nuclear bodies and chromatin
domains
C. A. Brackley,1 B. Liebchen,1 D. Michieletto,1 F. Mouvet,1 P. R. Cook,2 and D. Marenduzzo1
1SUPA, School of Physics and Astronomy, University of Edinburgh,
Peter Guthrie Road, Edinburgh, EH9 3FD, UK
2Sir William Dunn School of Pathology, University of Oxford, South Parks Road, Oxford, OX1 3RE, UK
Abstract
Fluorescence microscopy reveals that the contents of many (membrane-free) nuclear "bodies" exchange
rapidly with the soluble pool whilst the underlying structure persists; such observations await a satisfactory
biophysical explanation. To shed light on this, we perform large-scale Brownian dynamics simulations of a
chromatin fiber interacting with an ensemble of (multivalent) DNA-binding proteins; these proteins switch
between two states -- active (binding) and inactive (non-binding). This system provides a model for any
DNA-binding protein that can be modified post-translationally to change its affinity for DNA (e.g., like the
phosphorylation of a transcription factor). Due to this out-of-equilibrium process, proteins spontaneously
assemble into clusters of self-limiting size, as individual proteins in a cluster exchange with the soluble pool
with kinetics like those seen in photo-bleaching experiments. This behavior contrasts sharply with that ex-
hibited by "equilibrium", or non-switching, proteins that exist only in the binding state; when these bind to
DNA non-specifically, they form clusters that grow indefinitely in size. Our results point to post-translational
modification of chromatin-bridging proteins as a generic mechanism driving the self-assembly of highly dy-
namic, non-equilibrium, protein clusters with the properties of nuclear bodies. Such active modification also
reshapes intra-chromatin contacts to give networks resembling those seen in topologically-associating domains,
as switching markedly favors local (short-range) contacts over distant ones.
In all living organisms, from bacteria to man, DNA
and chromatin are invariably associated with binding
proteins, which organize their structure [1 -- 3]. Many of
these architectural proteins are molecular bridges that
can bind at two or more distinct DNA sites to form
loops. For example, bacterial DNA is looped and com-
pacted by the histone-like protein H-NS which has two
distinct DNA-binding domains [4]. In eukaryotes, com-
plexes of transcription factors and RNA polymerases sta-
bilize enhancer-promoter loops [5, 6, 6, 7], while HP1 [9],
histone H1 [10], and the polycomb-repressor complex
PRC1/2 [5, 11] organize inactive chromatin. Proteins
also bind to specific DNA sequences to form larger struc-
tures,
like nucleoli and the histone-locus, Cajal, and
promyeloleukemia bodies [13 -- 18]. The selective bind-
ing of molecular bridges to active and inactive regions
of chromatin has also been highlighted as one possible
mechanism underlying the formation of topologically as-
sociated domains (TADs) -- regions rich in local DNA
interactions [6, 6, 19].
From a biophysical perspective, a system made up of
DNA and DNA-binding proteins exhibits many kinds
of interesting and seemingly counter-intuitive behaviour,
such as the clustering of proteins in the absence of any at-
tractive interaction between them. This process is driven
by the "bridging-induced attraction" [20]. In conjunction
with the specific patterning of binding sites found on a
whole human chromosome in vivo, this attraction can
drive folding into TADs in the appropriate places on the
chromosome [6].
In the simple case where there is only a non-specific
DNA-protein interaction (i.e., proteins can bind to any
point on DNA), bridging-induced clustering can be un-
derstood as being due to a thermodynamic feedback loop:
binding of bridges to multiple DNA segments causes an
increase in local DNA concentration which, in turn, re-
cruits further DNA-binding proteins, and further itera-
tions then sustain the positive feedback. Subsequently,
the ensuing clusters coarsen, and eventually phase sepa-
rate into one macroscopic cluster of DNA-bound bridges
in equilibrium with a (diluted) pool of unbound pro-
teins [21, 22].
In the more complex case with specific
DNA-binding interactions, clustering is associated with
the formation of DNA loops. Looped structures incur
an entropic cost which increases superlinearly with the
number of loops, and can stop the growth of a cluster
beyond a critical size [6, 23 -- 26]. Such specific bind-
ing drives the formation of promoter-enhancer loops [2];
however there are several proteins which interact mainly
non-specifically with large regions of the genome, such
as histone H1 and other heterochromatin-associated pro-
teins [2]. For this latter class of proteins, the abun-
dance of binding sites in the nucleus would lead to clus-
ters that coarsen progressively. However, this indefinite
growth is not observed: we suggest that reversible post-
translational protein modifications may be the reason un-
derlying the arrested coarsening.
Specifically, here we consider a non-equilibrium bio-
chemical reaction which can modify DNA-binding pro-
teins. In our model, these proteins continuously switch
between an active, DNA-binding state, and an inactive,
non-binding, one. Such a reaction can arise in several
scenarios. For instance, a complex of transcription fac-
tors and an RNA polymerase might stabilize a promoter-
enhancer loop; upon transcription termination, the com-
plex could dissociate and the loop disappear [2, 3]. Al-
ternatively, phosphorylation, or other post-translational
modifications of transcription factors [27], may affect
1
their affinity for chromatin, as might a conformational
change in a protein or the reversible addition of a sub-
unit to a protein complex, which might be driven by ATP
hydrolysis.
In this work we show that introducing this non-
equilibrium mechanism strikingly broadens the range of
physical behaviour displayed by the chromatin/protein
ensemble. In particular, we find that including "switch-
ing" proteins which interact non-specifically with a chro-
matin fiber leads to qualitatively and quantitatively dif-
ferent results compared to "equilibrium" proteins (which
are always in the binding state). Switching (i.e., pro-
tein modification) arrests the coarsening triggered by
the bridging-induced attraction, and the size of the re-
sulting clusters can be tuned by altering the switching
rates. Furthermore, we show that if proteins bind both
specifically and non-specifically, switching results in the
formation of highly-dynamic clusters, which are qualita-
tively different from those formed by non-switching pro-
teins. In the former case, proteins in the cluster exchange
with the soluble pool, whilst the general shape of the
cluster persists. These dynamic clusters closely resem-
ble some of the nuclear bodies of eukaryotic cells. Fi-
nally, we consider a simplified model for the formation of
TADs in chromosomes, and show that protein switching
leaves the location of the domains unaltered, but strongly
disfavours long-range inter-TAD interactions. All these
findings point to an important and generic role of re-
versible protein modification in chromatin and nuclear
organization.
I. RESULTS
A. Protein switching arrests coarsening of chro-
matin bridges that bind non-specifically
We perform Brownian dynamics simulations of a flex-
ible chromatin fiber modeled as a bead-and-spring poly-
mer (thickness 30 nm, persistence length 90 nm) interact-
ing non-specifically with either non-switching or switch-
ing proteins. These proteins can bind to the fiber at more
than one location (in our case through a Lennard-Jones
potential; see Supporting Information for details on the
force field, and Fig. 1A,B for a schematic). For simplic-
ity, we assume a protein has the same size as the chro-
matin beads (a realistic assumption as each is likely to
be a protein complex). We also assume proteins stochas-
tically switch between binding and non-binding states at
an equal rate, α. [Relaxation of either of these assump-
tions does not qualitatively alter results.]
First, we consider the case of equilibrium, non-
switching proteins, where α = 0 (Fig. 1C). This case was
studied in [6, 20, 21, 28], and it was shown to lead to poly-
mer collapse [6, 21, 28] and clustering of bridges [20, 21]
depending on the protein concentration. For the concen-
trations used here, clusters coarsen and grow at the ex-
pense of smaller aggregates. During the early stages, this
2
Figure 1: Protein switching arrests coarsening. In (A-D), active
and inactive proteins are colored red and gray respectively; chromatin
is represented by strings of blue beads. (A) Schematic of the model
(Brownian dynamics simulations). (i) Proteins (lone spheres) switch
between red and gray states at rate α. (ii) Only proteins in the red
state can bind chromatin. (iii) Red and gray beads interact via steric
repulsion only. (iv) Proteins can bind to ≥ 2 sites to create molecular
bridges and loops. (B) Snapshots illustrating protein
binding/unbinding. Bound active proteins have clustered and
compacted chromatin. Bound active proteins 1 and 2 (gray circles)
switch and become inactive and dissociate (gray arrows); non-binding
sometimes bind to the cluster (red arrows). (C) Snapshots taken (i)
proteins a-c in the soluble pool (red circles) become active and
104 and (ii) 2 × 104 simulation units after equilibration. The
simulation involved a 5000-bead fiber (corresponding to 15 Mbp) and
N = 4000 non-switchable proteins, of which half are able to bind. (D)
As (C), but for N = 4000 switchable proteins (α = 0.0003 inverse
Brownian times). (E) Average cluster size as a function of time. Error
bars denote standard deviations of the mean. (i) Non-switching
proteins. (ii) Switching proteins; from top to bottom, α equals 0.0001,
0.0002, 0.0003, 0.0004 and 0.0005 inverse Brownian times.
resembles the Ostwald ripening characteristic of liquid-
gas phase separation; later on, we also observe coales-
cence of smaller clusters into larger ones (Movie S1). The
average cluster size -- measured as the number of bound
proteins per cluster -- increases with time with no sign
of saturation until all clusters merge into one (Fig. 1Cii,
and Movie S1). For early times, cluster size (which is also
proportional to its volume) increases approximately lin-
early with time, as would be expected for Ostwald ripen-
ing in density-conserving model B [2]. For later times,
cluster growth is much slower, with a sublinear exponent
(close to 0.25 for our parameters, see Fig. 1Ei). This
slowing is due to the underlying polymer dynamics -- as
in blob formation during the collapse of a homopolymer,
which is also slower than simple model B kinetics [4].
The dynamics with protein switching is remarkably dif-
ferent: now, coarsening is completely arrested, and the
system achieves a micro-phase separated state in which
clusters have a well-defined average size (Fig. 1D; Movie
S2). This size decreases with α (Fig. 1Eii). The ar-
rested phase separation can be understood intuitively as
follows. On the one hand, thermodynamics dictates that
the system should try to minimize interfaces, and this
leads to coarsening, initially via Ostwald ripening, giv-
ing the growth laws in Figure 1Ei. On the other hand,
non-equilibrium protein switching is a Poisson process,
so active proteins switch off at a constant rate α, and
leave the cluster.
[This is not the case for equilibrium
proteins which can only unbind thermodynamically; not
only is the unbinding rate slower, but such proteins are
also highly likely to rebind to a nearby site before ever
leaving a cluster]. Then, active proteins only have a
timescale of the order α−1 in which to form a cluster,
before a significant proportion of proteins in that cluster
inactivate. Hence, phase separation is arrested.
B. A mean field theory quantitatively explains the
arrest of coarsening, and predicts average cluster size
To understand more quantitatively how a non-
equilibrium biochemical reaction arrests coarsening, we
consider a simplified mean field theory which follows the
time evolution of the chromatin density ρ(, t), and of the
active protein density Φ(x, t). Our equations describe the
binding of the proteins to the chromatin together with
the diffusion of all components, and they read as follows,
ρ = Mρ∇2(cid:2)a1ρ − k∇2ρ − χΦ + gρ3(cid:3)
(1)
Φ = MΦ∇2 [a2Φ − χρ] − α(Φ − Φ0).
These equations can be formally derived starting from a
suitable underlying free energy density, and adding pro-
tein modification as a reaction term -- the details are
discussed in the Supporting Information.
In Eqs. (1),
Mρ and MΦ are the chromatin and protein mobility re-
spectively, so that Mρa1 ≡ D1 and MΦa2 ≡ D2 repre-
sent effective diffusion coefficients, while χ is the coef-
ficient describing bridging between active proteins and
chromatin. Further, g captures steric repulsion in the
chromatin fiber, κ accounts for effective surface tension
effects, and finally the last term in the equation for Φ
models the biochemical reaction, where proteins switch
from binding to non-binding, and vice versa, at a rate α.
For α = 0, Eqs. (1) ensure conservation of the global den-
sity of both chromatin and proteins -- in other words, this
is an example of generalized model B dynamics [2, 17].
To identify the key parameters in our system, we now
choose dimensionless time and space units tu = 1/α and
3
xu = (cid:112)D2/α and redefine Φ as Φ(Mρχ/D2). In these
units, our equations become
ρ = D0∇2ρ − A∇4ρ − ∇2φ + G∇2ρ3
Φ = ∇2Φ − X∇2ρ − (Φ − Φ0),
(2)
(3)
so that the whole parameter space is spanned by the
2); D0 =
four dimensionless numbers X = (χ2MρMΦ/D2
(D1/D2); A = αkMρ/(D2
2) and G = gMρ/D2.
A +
√D)2 where D = D0 + 3Gρ2
χ >(cid:112)kα/MΦ +(cid:112)(D2/MΦ)[D1/Mρ + 3gρ2
One solution of Eqs. (B6-B7) is given by the uniform
phase (ρ, Φ) = (ρ0, Φ0), which is stable in the absence of
bridging (χ = 0). To see how the interplay of bridging
and biochemical switching can create patterns, we per-
formed a linear stability analysis of this uniform state
(Fig. 2, detailed in the Supporting Information). The
√
result is that small perturbations of the uniform phase
grow if X > Xc = (
0.
This instability criterion translates in physical units to
0]. Thus, the
instability is driven by bridging, whereas diffusion of
chromatin and proteins, excluded volume, as well as pro-
tein modification, all tend to stabilize uniform chromatin-
protein distributions. Importantly, this bridging-induced
instability also works at very low protein concentration.
Calculating the wavenumber at the onset of instabil-
ity (see Supporting Information) unveils the remarkable
role played by the biochemical reaction for structure for-
mation. Specifically, we find qc = (D/A)(1/4) for the
dimensionless onset wavenumber, translating in physical
units to the following typical length scale,
(cid:20) D1D2 + 3Mρgρ2
0D2
(cid:21)1/4
Lc = 2π
αkMρ
.
(4)
Hence, in contrast to models without protein modifica-
tion, the present system exhibits a short wavelength in-
stability (Fig. 2), which turns into a long wavelength in-
stability only in the limit α → 0 (which would lead to
Lc → ∞, dotted black line in Fig. 2). Our linear sta-
bility analysis therefore suggests that the presence of the
biochemical reaction has qualitative consequences for the
clustering in the system, in that it leads to self-limiting
cluster sizes, or put differently, to micro-phase separation
rather than to macro-phase separation -- in full agreement
with the simulations shown in Fig. 1.
To further confirm that within our mean field theory
clusters cannot coarsen indefinitely, we also performed
a weakly non-linear expansion, through which we found
that the amplitude of the chromatin density fluctuations
close to the uniform state obey the "real Ginzburg Lan-
dau equation", which is associated with formation of sta-
tionary patterns of well-defined self-limiting size [3] (see
Supporting Information). Finally, Eq. (4) also predicts
that, at least close to the onset of clustering, the aver-
age number of proteins in any aggregate should scale as
c ∼ α−3/4, in good agreement with results from our
L3
simulations (Fig. S1).
are photobleached at a given time, and recovery of fluo-
rescence is then monitored (Fig. 3). The "fluorescence"
signal (proportional to the number of non-photobleached
active proteins in the clusters) recovers quickly in the
α (cid:54)= 0 case (Figs. 3Aiii, 3B), but not in the α = 0 case
(Fig. 3Aiv, 3Bi), at least for large values of the spe-
cific interaction strength. The dynamics of recovery can
be measured using the number of unbleached proteins
in the photobleached volume (Fig. 3Bi); this is propor-
tional to the fluorescence intensity measured in a stan-
dard FRAP experiment. Alternatively, the number of
unbleached proteins in clusters can be used (Fig. 3Bii).
Both approaches give similar recovery timescales, and
confirm that protein modification is required to create
clusters in which proteins can recycle.
The clusters found in Figure 3 typically contain ∼
20−100 proteins that recycle (Fig. S3A). Cluster size de-
pends on both protein concentration and interaction en-
ergy (e.g., in Fig. S3B there are only ∼ 5−10 proteins per
cluster). Therefore, this mechanism can produce clus-
ters with a wide range of sizes. Note that nuclear bodies
range from large nucleoli (up to several µm), through Ca-
jal, polycomb, and promyelocytic leukemia bodies (∼ 1
µm) [14 -- 16], to transcription factories containing ∼ 10
active transcription complexes (∼ 100 nm) [3, 34, 35].
Importantly, like most nuclear bodies, our clusters also
retain a "memory" of their shape. Thus, in Figure 3A,
when most of the components of the pink cluster on the
left have turned over, the general shape of the cluster
persists (see also Movies S4 and S5, and Fig. S4). This
is because the chromatin scaffold associated with the pro-
tein clusters (i.e., the sites of specific binding) retains a
general 3D structure that does not change much over time
(Fig. S5). Taken together, these results strongly support
the conjecture that nuclear bodies emerge from the aggre-
gation of bound switching proteins, and that switching
both arrests phase separation and ensures that bound
proteins continually exchange with the soluble pool.
Notably, the nuclear bodies which our clusters resem-
ble generally show FRAP recovery times in the range
of tens of minutes [36 -- 38]. These time scales are too
slow to be accounted for by diffusion, and too fast to be
compatible with the thermodynamic unbinding of tightly
bound proteins (see Supporting Information). Our re-
sults suggest an attractive alternative explanation: that
the recovery time, over which nuclear bodies recycle their
proteins, is instead linked to protein modification, and it
is simply proportional to α−1. Typical rates of post-
translational protein modification can be of the order of
several seconds (and will be slower within nuclear bodies
due to macromolecular crowding), and transcription ter-
mination occurs minutes after initiation. In light of this,
our simulations would predict recovery time scales of the
order of α−1, or minutes, broadly in agreement with those
measured experimentally [36 -- 38]. Further to this, there
is biological evidence that protein modifications can take
place within nuclear bodies [36]. For instance, enzymes
performing post-translational modifications are found in
Figure 2: Mean field theory predicts arrested coarsening with
protein modification. Dispersion relation, showing the growth rate,
fluctuations around the uniform solution of Eqs. (B6), for D = A = 1,
and X = 3.5 (cyan), corresponding to linear stability of the uniform
λ, as a function of the magnitude of the wavevector, q, for
phase, X = Xc = 4.0 (blue), marking the onset of instability, and
X = 4.5 (red), revealing the growth of clusters with a characteristic
length scale. The dotted black line shows a typical dispersion relation
in the absence of protein modification, which leads to a long
wavelength instability.
C. Switching proteins with specific binding self-
assemble into recycling nuclear bodies
The model considered in Figure 1 assumes proteins
bind non-specifically. While this is a good approximation
for generic heterochromatin-binding proteins in silenced
regions, most transcription factors bind to active regions
specifically and to most other DNA non-specifically [32].
Therefore, we consider proteins binding with high affinity
to every 20th bead (i.e., every 60 kbp), and with low affin-
ity to all others. [Similar results are expected for differ-
ent patterns of binding sites [6, 20].] Now bound proteins
self-assemble into clusters of self-limiting size even when
α = 0 (Fig. 3; Movie S3). In other words, coarsening is
always arrested. As suggested previously [20, 23, 25],
specific binding creates loops and loop clustering is asso-
ciated with entropic costs that scale super-linearly with
loop number, and this limits cluster growth [23].
Although coarsening is arrested whatever the value
of α, there is a major difference between the dynam-
ics of the equilibrium and switching proteins. Without
switching, proteins can only unbind thermodynamically,
which requires a long time: as a result, proteins rarely
exchange between clusters (Fig. 3, Movie S3). With
switching, there is a constant turnover of proteins within
the clusters, which recycle all their components over a
time ∼ α−1 (Fig. 3, Movie S4). Reducing the strength of
the specific interactions can also lead to protein turnover
(Fig. S2), but this requires fine tuning of the parame-
ters so as to simultaneously ensure stable binding and
the recycling of proteins in clusters. In contrast, protein
modification naturally leads to such recycling for any val-
ues of specific and non-specific binding affinity.
To quantitatively characterize
the dynamics of
turnover within clusters, we perform a simulated
fluorescence-recovery-after-photobleaching (FRAP) ex-
periment [33]. In such an experiment some of the clusters
4
Figure 3: In silico FRAP (Brownian dynamics simulations). (A) Snapshots taken 104 (i,ii) or 2 × 104 (iii,iv) after equilibration, during an
in silico FRAP experiment (only proteins -- and not chromatin beads -- are shown for clarity). (i) The simulation begins with N = 2000
equilibrium proteins, half of which are able to bind the chromatin fiber, both specifically (interaction strength 15 kB T , cut-off 1.8σ) to every
20-th bead in the polymer, and non-specifically (interaction strength 4 kB T , cut-off 1.8σ) to any other bead. After 104 time units, a structure
with multiple clusters forms. The snapshot shows only a portion of this, for clarity; 5 clusters of bound proteins have developed (unbound
proteins are colored gray, and bound ones in the 5 clusters a different color). Circled areas will be "photo-bleached". (ii) Photo-bleaching involves
making bound proteins invisible (the bleached proteins are still present in the simulation). (iii) If proteins can switch, clusters reappear in the
same general place (as new proteins replace their "bleached" counterparts). (iv) If proteins cannot switch (i.e., α = 0), clusters do not recover (as
their protein constituents do not recycle). (B) FRAP recovery. Error bars give SD of mean, and time is given in multiples of 104 simulation units;
the values of α, in units of inverse Brownian times, are as indicated in each panel. Only the post-bleaching signal is shown (the pre-bleaching
value would be constant and equal to 1 in these units). (i) Number of unbleached proteins in the bleached volume (a sphere of 50σ) as a function
of time, after bleaching. The signal is normalized with respect to the number of proteins initially in the bleached volume. (ii) Number of
unbleached proteins in clusters as a function of time after bleaching, after all proteins in clusters at a given instant are bleached. The signal is
normalized with respect to the proteins in clusters at the time of bleaching.
Cajal bodies [36], and phosphorylation or ubiquination
of the BMI1 subunit of the PcG PRC1 complex are im-
portant factors which determine the kinetics of exchange
in polycomb bodies [37].
D. Protein switching preserves TAD structure,
while suppressing long-range interactions
Clustering of bridging proteins can lead to the forma-
tion of chromatin domains [5, 6, 6, 20] resembling TADs
found in Hi-C data [19].
It is therefore of interest to
ask how switching affects TAD structure and dynamics.
Here, we return to a toy model first considered elsewhere:
the fiber has a regular pattern of binding and non-binding
regions (Fig. 4A), and each binding region spontaneously
and reproducibly assembles into a TAD which is flanked
by a disordered non-binding region [6]. The regular inter-
spersion of non-binding segments in Figure 4A fixes the
locations of TAD boundaries; consequently, clusters form
(Fig. 4B,C) at reproducible positions along the fiber, and
this -- in turn -- yields TADs seen in averaged contact
maps (Fig. 4D). Such patterns resemble those seen in
Hi-C data obtained from cell populations.
Variations in α have several effects (Fig. 4). First, the
configurations found at steady state are qualitatively dif-
ferent. Although cluster growth is limited for both α = 0
and α > 0, the (recycling) clusters formed by switching
proteins are much smaller (Figs. 4B, C, and Movie S6).
Second (and notwithstanding this qualitative difference),
the contact maps close to the diagonal are remarkably
similar (Fig. 4D; compare patterns on each side immedi-
ately next to the diagonal); this indicates that local TAD
structure is largely unperturbed by switching. However,
for α (cid:54)= 0 non-local (i.e., distant) contacts are strikingly
suppressed (Fig. 4D, compare patterns on each side far
5
from the diagonal), and higher order folding of one TAD
onto another is suppressed. This was demonstrated di-
rectly as follows. First, TADs were generated in the
presence of non-switching proteins, and then switching
turned on; the fraction of non-local contacts falls (Figs
4E, S6).
This behaviour can be explained as follows. First,
the time scale for the formation of TADs is compara-
ble to (or smaller than) that of protein recycling within
a TAD (see Supporting Information for an estimate of
such time scales). Computer simulations of TAD for-
mation in Drosophila and human chromosomes also sug-
gest that the local structure is formed very rapidly (at
most, in minutes) [5, 6]. Therefore it is plausible that
local TAD folding is fast enough not to be perturbed
much by protein modification. Second, when a particu-
lar protein switches from binding to non-binding, a con-
tact is lost, and it is likely that local ones can reform
faster than non-local ones. Furthermore, switching pro-
vides a non-equilibrium mechanism allowing faster large-
scale rearrangements, and so a more effective trimming
of entropically unfavourable long-ranged interactions. In
other words, active post-translational modification tilts
the balance in favour of local intra-TAD contacts at the
expenses of inter-TAD ones. This is consistent with
the sharp decay beyond the Mbp scale seen in Hi-C
data [19, 39].
II. CONCLUSIONS
To conclude, we have shown that active post-
translational protein modication (e.g., phosphoryla-
tion, methylation, acetylation [27], or any other non-
equilibrium reaction where a protein switches between
binding and non-binding states) has a profound and
generic effect on the behaviour of a chromatin-protein
mixture. The interplay between protein bridging and
protein modification is therefore an important principle
underlying nuclear organization within eukaryotes.
First, it was previously shown that non-switching pro-
teins able to bind non-specifically to chromatin to form
molecular bridges assemble into clusters which have a
natural tendency to coarsen [20, 21]. Here we show that
switching changes the behaviour; cluster growth is self-
limiting (Fig. 1) -- a phenomenon which be understood
via a simple mean field theory (Fig. 2). This theory
provides an example of arrested phase separation, and
it can explain why nuclear bodies do not progressively
enlarge [13 -- 16], and why neighbouring clumps of hete-
rochromatin -- whether detected using classical staining
and microscopy, or through inspection of Hi-C contact
maps [19, 40] -- rarely merge into one super-domain.
Second, when equilibrium (i.e., non-switching) pro-
teins bind specifically to cognates sites on the chromatin
fiber, they also cluster; however, specific binding is known
to arrest the coarsening [6, 20]. But in contrast to what is
seen in photobleaching experiments [36, 37], bound pro-
teins in the ensuing clusters exchange little with the solu-
6
Figure 4: Switching promotes intra-TAD contacts and
suppresses inter-TAD ones. (A) Overview. Simulations involved
N = 2000 non-switching (α = 0) or N = 2000 switching proteins
(α = 0.0001 inverse Brownian times); for α = 0 half of the proteins are
binding. In both cases, interaction energy and cut-off were 4 kB T and
1.8σ. The fiber (length 15 Mbp) consisted of regularly-interspersed
segments containing runs of binding (blue) and non-binding (black)
beads (segment sizes 900 and 300 kbp, respectively). (B,C) Snapshots
taken after 105 simulation units. Non-binding and binding proteins
are colored gray and red, respectively. (D) Contact maps (averages
from 10 simulations) for non-switching (top-left triangle), and
switching proteins (lower-right triangle). The scale (right) indicates
contact frequencies. (E) The evolution of the ratio of non-local
contacts over time. A local (non-local) contact is one between beads
separated by less (more) than 1.2 Mbp along the fiber. Here the
simulation was run for 105 simulation units with non-switching
proteins; switching was then turned on (α = 0.0001 inverse Brownian
times) and the simulation was run for a further 105 simulation units.
ble pool. Moreover, the time scales seen in such bleaching
experiments are too slow to be accounted for by diffusion,
and too fast to be compatible with the thermodynamic
unbinding of tightly-bound proteins. Protein modifica-
tion provides a neat solution to this paradox: dynamic
clusters naturally emerge during simulations, with con-
stituent proteins recycling on a time-scale proportional
to the inverse switching rate, α−1 (Fig. 3). Importantly,
when clusters in simulations are "photobleached", they
behave like nuclear bodies seen in vivo -- they retain a
"memory" of their shape, despite the continual exchange
with the soluble pool.
Finally, switching affects large-scale chromatin orga-
nization. Bridging-induced clusters are associated with
the formation of chromatin domains, reminiscent of the
TADs observed in Hi-C data [19]. Using a fiber patterned
in such a way that it spontaneously folds into TADs, we
find that switching has little effect on local TAD orga-
nization, but strongly suppresses inter-TAD interactions;
local contacts are favored over non-local ones (Fig. 4).
We expect that similar trends should be observed in more
complex models for bridging-induced chromosome orga-
nization, such as those in Refs. [6, 6, 20].
While here we focus on a flexible chromatin fiber, we
expect that near-identical results should be reached with
a semi-flexible one [20, 22]; then, our conclusions should
also apply to bacterial DNA. We also expect similar re-
sults to be obtained with more complex pathways be-
tween active and inactive states (e.g., modeling the cyclic
flooding of proteins into nuclei, or their cyclic synthe-
sis/degradation) and it would be of interest to investigate
these scenarios.
In summary, we demonstrated how non-equilibrium
processes involving ephemeral protein states can provide
a simple way of understanding how dynamic nuclear bod-
ies of self-limiting size might form, and how chromosomal
domains at the larger scale might be organized.
This work was funded by ERC (Consolidator Grant
648050, THREEDCELLPHYSICS), and by a Marie
Skodowska-Curie Intra European Fellowship (G. A. No.
654908).
[1] Calladine, C. R., Drew, H. R., Luisi, B. F., Travers, A. A.
(2004). Understanding DNA, 3rd Edition, Elsevier Aca-
demic Press, London.
[2] Alberts, B. et al. (2002). Molecular biology of the cell, Gar-
land Science, New York.
[3] Cook, P. R. (2001). Principles of Nuclear Structure and
Function, Wiley-Liss (New York) (2001).
[4] Dame, R. T., M. C. Noom, M. C., Wuite, G. J. (2006).
Bacterial chromatin organization by H-NS protein unrav-
elled using dual DNA manipulation. Nature 444, 387-390.
[5] Simonis, M. et al. (2006) Nuclear organization of active
and inactive chromatin domains uncovered by chromo-
some conformation capture-on-chip (4C). Nat. Genet. 38,
1348-1354.
[6] Barbieri, M. et al. (2012). Complexity of chromatin folding
is captured by the strings and binders switch model. Proc.
Natl. Acad. Sci. USA 109, 16173-16178.
[7] Brackley, C. A. et al. (2016). Predicting the three-
dimensional
folding of cis-regulatory regions in mam-
malian genomes using bioinformatic data and polymer
models. Gen. Biol. 17, 59 (2016).
[8] Brackley, C. A., Johnson, J., Kelly, S., Cook, P. R., Maren-
duzzo, D. (2016). Simulated binding of transcription fac-
tors to active and inactive regions folds human chromo-
somes into loops, rosettes and topological domains. Nucl.
Acids Res. 44, 3503-3512.
[9] Kilic, S., Bachmann, A. L., Bryan, L. C., Fierz, B. (2015).
Multivalency governs HP1α association dynamics with the
silent chromatin state. Nat. Commun. 6, 7313 (2015).
[10] Harshman, S. W., Young, N. L., Parthun, M. R., Fre-
itas, M. A. (2013). H1 histones: current perspectives and
challenges. Nucl. Acids Res. 41, 9593-9609 (2013).
[11] Wani, A. H. et al. (2016). Chromatin topology is coupled
to Polycomb group protein subnuclear organization. Nat.
Comm. 7, 10291.
[12] Michieletto, D., Marenduzzo, D., Wani, A. H. (2016)
folding path-
interphase chromosomes.
Chromosome-wide simulations uncover
way and 3D organization of
arXiv:1604.03041.
[13] Zhu, L., Brangwynne, C. P. (2015). Nuclear bodies: the
emerging biophysics of nucleoplasmic phases. Curr. Opin.
Cell Biol. 34, 23-30.
[14] Sleeman, J.E., Trinkle-Mulcahy, L. (2014). Nuclear bod-
ies: new insights into assembly/dynamics and disease rel-
evance. Curr. Opin. Cell Biol. 28, 76-83.
[15] Mao, Y. S., Zhang, B., Spector, D. L. (2011) Biogenesis
and function of nuclear bodies. Trends Genet. 27, 295-306.
[16] Pirrotta, V., Li, H. B. (2012) A view of nuclear polycomb
bodies. Curr. Opin. Gen. Devel. 22, 101-109.
[17] Berry, J., Weber, S. C., Vaidya, N., Haataja, M., Brang-
wynne, C. P. (2015). RNA transcription modulates phase
transition-driven nuclear body assembly.
[18] Brangwynne C. P., Tompa P., Pappu R. V., (2015) Poly-
mer physics of intracellular phase transitions. Nat Phys.
11, 11, 899-904.
[19] Dixon, J. R., Selvaraj, S., Yue, F., Kim, A., Li, Y., Shen,
Y., Hu, H., Liu, J. S., Ren, B. (2012). Topological domains
in mammalian genomes identified by analysis of chromatin
interactions. Nature 485, (2012) 376-380.
[20] Brackley, C. A, Taylor, S., Papantonis, A., Cook, P. R.,
Marenduzzo, D. (2013). Nonspecific bridging-induced at-
traction drives clustering of DNA-binding proteins and
genome organization. Proc. Natl. Acad. Sci. USA 110,
E3605-3611.
[21] Johnson, J., Brackley, C. A., Cook, P. R., Marenduzzo,
D. (2015). A simple model for DNA bridging proteins and
bacterial or human genomes: bridging-induced attraction
and genome compaction. J. Phys. Cond. Matt. 27, 064119.
[22] Le Treut, G., Kepes, F., Orland, H. (2016). Phase Be-
havior of DNA in the Presence of DNA-Binding Proteins
Biophys. J. 110, 51-62 (2016).
[23] Marenduzzo, D., Orlandini, E. (2009). Topological and
entropic repulsion in biopolymers. JSTAT, L09002.
[24] Cates, M. E., Witten, T. A. (1986). Chain conformation
and solubility of associating polymers. Macromolecules 19,
732-739.
[25] Orlandini, E., Garel, T. (1998). Collapse transitions of a
periodic hydrophilic hydrophobic chain, Eur. Phys. J. B
6, 101-110.
[26] Scolari, V. F., Lagomarsino, M. C. (2015). Combined col-
lapse by bridging and self-adhesion in a prototypical poly-
mer model inspired by the bacterial nucleoid. Soft Matter
11, 1677-1687.
[27] Tootle, T. L., Rebay, I. (2005). Post-translational modifi-
cations influence transcription factor activity: a view from
the ETS superfamily. Bioessays 27, 285-298 (2005).
[28] Nicodemi, M., Prisco, A. (2009). Thermodynamic Path-
ways to Genome Spatial Organization in the Cell Nucleus.
Biophys. J. 96, 2168-2177.
[29] Chaikin, P. M., Lubensky, T. C. (2000). Principles of
Condensed Matter Physics, Cambridge University Press.
[30] Byrne, A., Kiernan, P., Green, D., Dawson, K. A. (1995).
Kinetics of homopolymer collapse. J. Chem. Phys. 1012,
573-577 (1995).
[31] Cross, M. C., Hohenberg, P. C. Pattern formation outside
of equilibrium. Rev. Mod. Phys 65, 851-1112 (1993).
[32] Sheinman, M., Benichou, O., Kafri, Y., Voituriez, R.
(2012). Classes of fast and specific search mechanisms for
7
proteins on DNA. Rep. Progr. Phys. 7, 026601.
[33] Mueller, F., Mazza, D., Stasevich, T. J., McNally, J. G.
(2010).FRAP and kinetic modeling in the analysis of nu-
clear protein dynamics: what do we really know? Curr
Opin Cell Biol. 22, 403-411.
[34] Papantonis, A., Cook, P.R. (2013). Transcription facto-
ries; genome organization and gene regulation. Chem. Rev.
113, 8683-8705.
[35] Marenduzzo, D., Micheletti, C., Cook, P. R. (2006).
Entropy-driven genome organization. Biophys. J. 90,
3712-3721.
[36] Handwerger, K. E., Murphy, C., Gall, J. G. (2003)
Steady-state dynamics of Cajal body components in the
Xenopus germinal vesicle. J. Cell. Biol. 160, 495-504.
[37] Hern´andez-Munoz I, Taghavi P, Kuijl C, Neefjes J, van
Lohuizen M. (2005) Association of BMI1 with polycomb
bodies is dynamic and requires PRC2/EZH2 and the
maintenance DNA methyltransferase DNMT1. Mol Cell
Biol. 25 11047-58.
[38] Kimura, H., Sugaya, K., Cook, P. R. (2002) The tran-
scription cycle of RNA polymerase II in living cells. J.
Cell Biol. 159, 777-782.
[39] Sanborn, A. L. et al. (2015). Chromatin extrusion ex-
plains key features of loop and domain formation in wild-
type and engineered genomes. Proc. Natl. Acad. Sci. USA
112, E6456-E6465
[40] Sexton, T. et al. (2012). Three-dimensional
folding
and functional organization principles of the Drosophila
genome. Cell 148, 458-472.
8
Supplementary Information
Here we give more details on the simulations (includ-
ing parameter values), and on the continuum mean field
model (derivation, linear stability analysis and amplitude
equation); we also show additional results and figures
which are discussed in the main text.
Appendix A: Simulation Details
The chromatin fiber is modeled as a bead-spring poly-
mer with finitely-extensible non-linear elastic springs via
a Kremer-Grest model [1]. To map length scales from
simulation to physical units, we can, e.g., set the diam-
eter, σ, of each bead to ∼ 30nm(cid:39) 3 kbp (assuming an
underlying 30 nm fiber; of course, all our results would
remain valid with a different mapping).
Letting ri and di,j ≡ rj−ri be respectively the position
of the centre of the i-th bead and the vector of length
di,j between beads i and j, we can express the potential
modeling the connectivity of the chain as
UFENE(i, i + 1) = − k
2
R2
0 ln
1 −
(cid:34)
(cid:18) di,i+1
(cid:19)2(cid:35)
,
R0
for di,i+1 < R0 and UFENE(i, i + 1) = ∞, otherwise; here
we chose R0 = 1.6 σ and k = 30 /σ2.
The bending rigidity of the chain is described through
a standard Kratky-Porod potential, as follows
(cid:20)
1 − di,i+1 · di+1,i+2
di,i+1di+1,i+2
(cid:21)
,
Ub(i, i + 1, i + 2) =
kBT lp
σ
where we set the persistence length lp = 3σ (cid:39) 90 nm,
which is reasonable for a chromatin fiber.
The steric interaction between a chromatin bead, a,
and a protein bridge, b (of sizes σa = σb = σ), is modeled
through a truncated and shifted Lennard-Jones potential
(cid:34)(cid:18) σ
di,j
(cid:19)12 −
(cid:18) σ
(cid:19)6 −
(cid:18) σ
(cid:19)12
+
di,j
rc
(cid:18) σ
rc
ULJ(i, j) = 4ab
(cid:19)6(cid:35)
for di,j < rc and 0 otherwise. This parameter, rc, is
the interaction cutoff; it is set to rc = 21/6σ for inactive
proteins, in order to model purely repulsive interactions,
and to rc = 1.8σ for an active protein, so as to include
attractive interactions.
In both cases, the potential is
shifted to zero at the cut-off in order to have a smooth
curve and avoid singularities in the forces. Purely re-
pulsive interactions, such as those between inactive pro-
teins and chromatin segments, are modeled by setting
ab = kBT , while attractive interactions are modeled us-
ing: (i) ab = 3kBT (for non-specific interactions, Fig. 1);
(ii) ab = 15kBT and ab = 4kBT (for non-specific and
specific interactions respectively, Fig. 3); (iii) = 4kBT
(for non-specific interactions, Fig. 4); or (iv) as specified
in Supporting Figure captions in other cases.
9
The total potential energy experienced by bead i is
given by
Ui =
+
(cid:88)
(cid:88)
j
(cid:88)
j
k
UFENE(i, j)δj,i+1+
Ub(i, j, k)δj,i+1δk,i+2 +
(A1)
ULJ(i, j),
(cid:88)
j
and its dynamics can be described by the Langevin equa-
tion
mri = −ξ ri − ∇Ui + ηi,
(A2)
where m is the bead mass, ξ is the friction coefficient, and
ηi is a stochastic delta-correlated noise. The variance of
each Cartesian component of the noise, σ2
η, satisfies the
usual fluctuation dissipation relation σ2
the LJ time τLJ = σ(cid:112)m/ and the Brownian time
As is customary [1], we set m/ξ = τLJ = τB, with
η = 2ξkBT .
τB = σ/Db, where is the simulation energy unit, equal
to kBT , and Db = kBT /ξ is the diffusion coefficient of a
bead of size σ. From the Stokes friction coefficient for
spherical beads of diameter σ we have that ξ = 3πηsolσ
where ηsol is the solution viscosity. One can map this
to physical units by setting the viscosity to that of the
nucleoplasm, which ranges between 10 − 100 cP, and by
setting T = 300 K and σ = 30 nm, as above. From this
it follows that τLJ = τB = 3πηsolσ3/ (cid:39) 0.6 − 6 ms; τB
is our time simulation unit, used when measuring time
in the figures in the main text and in this Supporting
Information. The numerical integration of Eq. (A2) is
performed using a standard velocity-Verlet algorithm
with time step ∆t = 0.01τB and is implemented in the
LAMMPS engine. We perform simulations for up to
2×105 τB, which correspond to 2-20 minutes in real time.
Appendix B: Phenomenological Mean Field Model
for Bridges with Active Modification
,
In our particle based simulations we observed the
growth of clusters due to bridging interactions (see main
text). When protein activation-inactivation reactions
were absent, these clusters coarsened, resulting in one
large macroscopic cluster in steady state. However, in the
presence of these reactions, the clusters coarsened only
up to a self-limiting size. To better understand this tran-
sition from macrophase separation to microphase sepa-
ration, and the involved length scales, we now develop
a phenomenological minimal model for the dynamics of
chromatin and proteins. We describe the distribution of
chromatin via the probability density field ρ(x, t), and
the density of active, or binding, and inactive, or non-
binding, proteins by Φa(x, t) ≡ Φ(x, t) and Φi(x, t) re-
spectively.
The starting point for our model is the free energy
F =(cid:82) f (x)dx where f is the free energy density:
f =
D(cid:48)
1
2
ρ2 +
D(cid:48)
2
2
Φ2 − χ(cid:48)ρΦ +
k(cid:48)
2
(∇ρ)2 +
g(cid:48)
4
ρ4.
(B1)
Here, the first two terms describe diffusion of chromatin
and proteins respectively, the third term describes the
energy gain through bridging and the last two terms,
multiplied by k(cid:48), g(cid:48), respectively penalize sharp interfaces
due to interfacial tension, and strong accumulations of
chromatin due to short ranged repulsions.
Assuming diffusive dynamics here and using the fact
that in the absence of protein modification, the number
density of all species (ρ, Φ, Φi) is conserved, we can derive
the equations of motions for our fields as done for model
B dynamics [2]. However, in the presence of active pro-
tein modification, we need an additional reaction term,
so that our equations of motion read
Appendix C: Linear Stability Analysis
To better understand in which parameter regimes we
should expect (i) a uniform distribution of chromatin
and proteins, (ii) cluster growth proceeding to macro-
scopic phase separation and (iii) microphase separation,
we now perform a linear stability analysis. This analysis
will equip us with a prediction for the self-limiting cluster
size in regime (iii), matching the results of our particle
based simulations. We therefore study the response of the
uniform phase to small perturbations in the density fields
(ρ, Φ). Linearising Eqs. (B6,B7) around the uniform so-
lution (ρ, Φ) = (ρ0, Φ0), where ρ0 is the DNA density as
fixed by the initial state, leads to the following equations
of motion for the fluctuations ρ(cid:48) = ρ − ρ0, Φ(cid:48) = Φ − Φ0:
ρ(cid:48) = D∇2ρ(cid:48) − A∇4ρ − ∇2Φ(cid:48),
Φ(cid:48) = ∇2Φ(cid:48) − X∇2ρ(cid:48) − Φ(cid:48).
(C1)
(C2)
ρ = −Mρ∇2 δF
Φa = −Ma∇2 δF
δρ
,
δΦa
− αΦa + βΦi.
(B2)
(B3)
Here, we defined D := D0 + 3Gρ2
0. Fourier transforming
Eqs. (C1,C2) and using Q := q2 leads to the following
dispersion relation (or characteristic polynomial),
Here Mρ and Ma are dimensionless mobility coefficients
of chromatin and activated proteins respectively, while α
and β are the activation and inactivation rates for pro-
teins. Since inactive proteins do not bind, we assume
that they diffuse quickly, i.e. that their density field is
uniform.
Now integrating Eq. (B3) over the whole system and
denoting the total number of active and inactive pro-
teins with Na(t) and Ni(t) respectively, we obtain Na =
−αNa + βNi. Conservation of the total protein num-
Ni = (1 + β/α)Ni
ber N = Na + Nb now yields
which approaches the steady state Ni = αN/(α + β),
i.e. Φi = α/(α + β), exponentially fast. Now defining
Φ0 := (β/α)Φi = β/(α + β) (and ignoring short-time
effects due to possible 'imbalances' between active and
inactive proteins in the initial state), Eqs. (B2,B3) re-
duce to:
ρ = Mρ∇2[a1ρ − k∇2ρ − χΦ + gρ3],
Φ = MΦ∇2[a2Φ − χρ] − α(Φ − Φ0),
(B4)
(B5)
where for simplicity hereon we drop the subscript a on
Φa for active proteins. We also introduced D1 = Mρa1
and D2 = MΦa2.
To further reduce these equations and to identify a
minimal set of dimensionless control parameters, we now
choose time and space units tu = 1/α and xu =(cid:112)D2/α
and redefine Φ = ΦχMρ/D2. This leads to
ρ = D0∇2ρ − A∇4ρ − ∇2φ + G∇2ρ3,
Φ = ∇2Φ − X∇2ρ − (Φ − Φ0).
(B6)
(B7)
1
2
(cid:113)
[−1 − Q (1 + D + AQ) ±
[−1 + Q(D − 1 + AQ)]2 + 4Q2X
(cid:21)
,
(C3)
λ(Q) =
±
which links the growth rate λ of the fluctuation with its
wavevector Q. An analysis of this relation leads us to the
instability criterion
(cid:112)
√
X >
XC :=
A +
√
√
D,
which translates, in physical units, to
(cid:114) kα
χ >
+
MΦ
(cid:115)
(cid:20) D1
Mρ
D2
MΦ
+ 3gρ2
0
(C4)
(C5)
(cid:21)
.
This criterion determines the transition line (hypersur-
face) between regions (i) and (ii/iii) in the parameter
space. Hence, if the bridging interactions are sufficiently
large, small fluctuations around the uniform state will
grow to form clusters. Remarkably, this instability and
the corresponding emergence of order (clustered phase) is
not contingent on the presence of a certain minimal pro-
tein (or DNA) density, suggesting that even a very low
protein concentration is sufficient to trigger clustering.
aration to microseparation (at the onset of
ity),
instability first occurs.
To map out the transition line from macrophase sep-
instabil-
it is useful to consider the wavelength at which
From Eq. (C3) and qc =
D)2, we find qc = (D/A)1/4, corre-
∂qλ(q) = 0X=(
√
A+
√
sponding, in physical units, to the length scale
(cid:18) D1D2 + 3Mρgρ2
0D2
(cid:19)1/4
αkMρ
.
(C6)
That is, our parameter space is spanned by the four
2); D0 =
dimensionless numbers X = (χ2MρMΦ/D2
(D1/D2); A = αkMρ/(D2
2) and G = gMρ/D2.
Lc =
2π
qc
= 2π
10
(cid:18)0
(cid:19)
0
(cid:18)∂2
We begin by rewriting Eqs. (B2,B3) as
(cid:19)
(cid:18) ρ(cid:48)
Φ(cid:48)
L
+ N −
(cid:18) ρ(cid:48)
(cid:19)
Φ(cid:48)
=
,
(D1)
where the linear operator L and the nonlinear term N
represent
(cid:18)D∂2
(cid:19)
x −∂2
x − 1
∂2
x
x − A∂4
−X∂2
x
L =
; N = G
xρ(cid:48)3 + 3ρ0∂2
xρ(cid:48)2
0
(D2)
(cid:19)
.
Now, we replace (as usual, see [3])
√
X → (1 + )XC;
where = (X − XC)/XC and expand the fields as
∂x → ∂x +
∂X ;
∂t → ∂T (D3)
∞(cid:88)
∞(cid:88)
ρ(cid:48) =
n/2ρn−1; Φ(cid:48) =
n/2Φn−1.
(D4)
n=1
n=1
Next, we plug these expansions into Eqs. (D2) and
solve the resulting equations to lowest order (1/2). Us-
ing the Ansatz ρ0 = A exp (iqcx) + c.c. and Φ0 =
Aφ exp (iqcx) + c.c. with amplitudes A,Aφ, we find qc =
(D/A)1/4 reproducing the corresponding result from our
linear stability analysis (see above), as well as Aφ =
c ) = A√DXC which fixes the relation be-
A(D + Aq2
tween the amplitudes of both density fields. The solution
of our perturbative equations to order 1/2 then reads
ρ(cid:48) = 2A cos qcx with the so-far unknown amplitude A.
The result to order turns out not to be particular use-
ful for our purpose, as solving it would provide us with
a similar result as to order 1/2), but with another un-
known amplitude A(cid:48) yielding a higher order correction to
the solution ρ(cid:48) = 2A cos qcx. Since we are looking only
for the lowest order result in we directly consider the
perturbative equations of motion to order 3/2. As usual
[3], we do not attempt to solve the corresponding equa-
tions explicitly, but apply Fredholm's theorem providing
solvability conditions, which determine an equation of
motion for A. After a long but straightforward calcula-
tion and transforming back to coordinates t, x we find:
ct A = A + cx∂2
xA + c3A3,
(cid:19)
(cid:18)
1
D
,
XC
(cid:114) A
4A√DXC
3G√DXC
1 +
,
.
where
ct =
cx =
c3 =
(D5)
(D6)
(D7)
(D8)
Eq.
(D5) is a variant of the real Ginzburg-Landau
equation, here describing, together with the coefficients
Eqs. (D6 -- D8), the dynamics of chromatin and proteins
close to the onset of instability. In this equation /(cttu)
Figure S1: Comparision between mean field predictions and
simulation results for cluster size. Simulations are like the ones
illustrated in Fig. 1 of the main text (except for the value of α, which
is varied). Dots show saturation values (± SD) of number of particles
per cluster N (after 1.5 × 105 simulation units); the line shows a
least-squares fit with a slope of −0.756.
Thus, in an infinite system, coarsening only occurs for
α = 0.
[In finite systems macrophase separation is ob-
served if α is small enough that Lc exceeds the system
size.] From this analysis we expect the average parti-
c ∝ α−3/4 (at
cle number per cluster to scale as N ∝ L3
least close to the onset of instability). This value agrees
well with the numerically observed scaling of N ∝ α−0.76
(Fig. S1), supporting the view that the essential physics
of chromatin clustering can be described and understood
within our simplified mean field theory.
For completeness, we also calculate the boundaries of
the instability band from Eq. (C3), which, after translat-
ing back into physical units (for M1 = M2 = 1), read as
follows:
(cid:114)
ν ±(cid:113)
ν = (cid:2)χ2 − D1D2 − 3D2gρ2
1√
2D2K
q± =
0 − Kα(cid:3) .
(C8)
At the onset of instability, we find ν → 4D2Kα(D1 +
3gρ0) and hence we recover the α1/4-scaling of the onset
mode. In contrast, the boundaries of the instability band
scale in a more complicated way which is nonuniversal in
α.
ν2 − 4D2Kα(D1 + 3gρ2
0),(C7)
Appendix D: Amplitude Equations
We now perform a perturbative analysis of the linearly
unstable modes (fluctuations) close to onset of instability.
This analysis will lead us to a further reduced effective
model, describing the linear growth and nonlinear satu-
ration of chromatin clusters on large scales and at long
timescales.
11
(cid:112)/c3 de-
√
(cid:104)√
√D(cid:105)2
is the initial growth rate of protein clusters; xu
scribes the amplitude of their saturation (related to their
density) for a given X >
cx is a
correlation length, describing a scale of spatial modula-
tions of the saturation amplitude of DNA clusters.
and xu
A +
Although we equipped our original equilibrium model
with reaction terms which drive it out of equilibrium, its
large scale and long time dynamics (i.e., Eq. (D5)) can
be effectively mapped (at least close to onset of instabil-
ity) onto a potential system with the following Lyapunov
functional:
V[A] =
(cid:104)−A2 +
x∂XA2(cid:105)
A4 + c2
(cid:90)
,
(D9)
dx
c3
2
A =
−1
cT
δV
δA .
(D10)
Hence, quite remarkably, the dynamics of the present
reaction-diffusion system can be mapped, within this ap-
proximation, onto a system which is purely relaxational.
shows recovery also for equilibrium bridges, if the specific
and non-specific interactions are carefully tuned. How-
ever, protein modification provides a more robust way
to achieve this, which simultaneously allows stable bind-
ing (when the protein is in the active state), and fast
turnover (due to the unbinding and diffusion of inactive
proteins).
Figure S3 shows the cluster size for different parame-
ter values for the case of non-specific protein-chromatin
interactions. This demonstrates that it can be varied
significantly (by about an order of magnitude), and is
particularly sensitive to the protein concentration.
(A)
(B)
Figure S2: Comparison of FRAP recovery for non-switching
and switching proteins. FRAP recovery, measured as the number
of unbleached proteins which are in the bleached volume after
bleaching. The signals are normalized with the number of proteins in
the bleached volume at the time of bleaching. As in Fig. 3Bi, the
bleached volume is a sphere of size 50σ. Error bars give SD of mean,
and time is given in multiples of 104 simulation units. Values of the
specific and non-specific interactions, and of α, were respectively:
15kB T , 4kB T , 0 (red curve), 8kB T , 3kB T , 0 (green curve), and
15kB T , 4kB T , 0.0001 inverse Brownian times (blue curve). It can be
seen that varying the values of non-specific and specific interactions
can lead to FRAP recovery also for α = 0 (green curve), although, in
the absence of fine tuning, this is to a smaller extent with respect to
α (cid:54)= 0 (blue curve).
Figure S3: Cluster size with specific binding. (A) Plot of the
average number of proteins in a cluster versus time (± SD), for
N = 2000 switching proteins binding to the chromatin fiber, both
specifically (interaction strength 15 kB T , cut-off 1.8σ), to every 20-th
bead in the polymer, and non-specifically (interaction strength 4
kB T , cut-off 1.8σ) to any other bead. From top to bottom, curves
correspond to α = 0 (in which case half of the proteins are
non-binding, and half binding), 0.0001, 0.0002, 0.0003, 0.0004, 0.0005
respectively. (B) Same as (A), but now for N = 500 switching
proteins, with specific interaction strength of 8kB T and non-specific
interaction of 3kB T ; the interaction cut-off is 1.8σ.
Appendix E: Additional simulation results
In this section we present additional simulation results,
which complement those discussed in the main text.
Figure S2 shows that the FRAP signal (following sim-
ulated photobleaching of a spherical spot of size 50 σ)
Figs. S4 and S5 highlight some further properties of
the recycling clusters. In particular, Fig. S4 shows that
these clusters retain memory of their shape even as the
proteins which constitute them change. Figure S5 shows
the dynamics of some protein and chromatin beads with
and without modification. Without modification, once
proteins bind to a cluster they diffuse little for the rest
12
(Ai)
(Bi)
(Aii)
(Bii)
Figure S4: Switching proteins form clusters which retain memory of their shape. This figure follows the evolution of clusters in a
simulation analogous to that of Fig. 3A in the main text; the same parameters apply. Only proteins -- and not chromatin beads -- are shown for
clarity. (A) Snapshots taken 104 time units after equilibration, for non-switching proteins, showing two clusters (beads are colored according to
the cluster they belong to); (ii) shows another cluster. (B) Snapshots of the same regions shown in (A) after another 105 simulation units, and
after allowing the proteins to now switch (α = 0.0001τ
−1
B ). Clusters recycle their constituent proteins whilst retaining a very similar shape.
of the simulation, whereas with modification they sample
the whole simulation domain. Contrary to this, the dy-
namics of the chromatin beads within a cluster is similar
with and without modification: they diffuse very little.
This explains why clusters keep their shape: while pro-
teins bind and unbind, the underlying chromatin back-
bone is largely unchanged.
Finally, Fig.
S6 shows how the effect of protein
switching on the ratio between non-local and local
contacts, shown in Fig. 4 in the main text, is affected
by the values of non-specific and specific interactions.
13
Appendix F: Timescale estimates for FRAP exper-
iments and TAD folding
Here we provide a series of simple estimates for the
value of the relevant timescales in our problems. Con-
sider first a fluorescence-recovery-after-photobleaching,
or FRAP, experiment, where a cluster of size σcl ∼ 0.1−1
µm is inside the bleached spot, which we imagine has a di-
ameter of σFRAP ∼ 1 µm. In this Section, as previously,
σ will instead denote the size of a typical chromatin-
binding complex, or chromatin bead (as previously, we
imagine this is ∼ 30 nm).
What is the timescale for the recovery of the FRAP
signal? Clearly, this depends on the underlying dynamics
of the bleached/unbleached proteins. If proteins diffuse
freely, then unbleached proteins can enter the bleach spot
to give recovery within a time, τdiff , proportional to
τdiff ∼ σ2
FRAP
D
.
(F1)
For a protein size σ ∼ 30 nm, and if the nucleoplasm
viscosity is 10 cP (ten times that of water), the diffusion
coefficient is ∼ 1.4µm2 s−1, so that τdiff ∼ 1s, which is
too fast to account for FRAP response of nuclear bodies
(furthermore, of course, freely diffusing proteins could
not self-organise into clusters).
If,
instead, non-switching binding proteins create a
cluster, then the FRAP signal recovers when some pro-
teins unbind, and others replace these from the soluble
(unbleached) pool. As the former process is slower than
the latter (which relies again on diffusion), we can equate
the FRAP recovery timescale to the time needed for an
equilibrium protein to unbind from the cluster, which can
be estimated as,
(cid:18) ∆U
(cid:19)
kBT
τnon−switch ∼ σ2
cl
D
exp
(F2)
Figure S5: Trajectories of proteins and high-affinity chromatin
beads. Simulations are as in Fig. 3 of the main text; the same
parameters apply. Positions of proteins and chromatin beads are
shown in a 2D projection of the simulation domain, positions on the
axes are measured in units of σ. (A) Non-switching proteins. (i) Red,
proteins, recorded every 100 τB in a simulation (total length 1.5 × 105
simulation units. In this case, all three proteins remain bound to one
green and blue circles denote positions of three non-switching
cluster throughout the time series. (ii) Red, green and blue circles
denote positions of three high affinity chromatin beads, again
recorded every 100 τB in the same simulation. All three chromatin
beads remain in the same cluster. (B) Same as (A), but for switching
proteins (α = 0.0001 inverse Brownian times). Now the three
switching proteins diffuse through the whole space, while the three
chromatin beads are still confined; this shows that the underlying
scaffold of the cluster persists as the proteins are recycled.
where ∆U indicates the strength of chromatin-protein
interaction. If we assume an interaction of 10 kcal/mol,
consistent with either multiple non-specific or a single
specific DNA-protein interactions, then τnon−switch > 105
s, which is too slow to account for the FRAP recovery
observed in nuclear bodies. Clearly, changing ∆U will
change τnon−switch, but in order for the estimate to be
in the observed range, the interaction energy would have
to be finely tuned, and would be significantly lower than
that seen in typical DNA-protein interactions.
If, finally, switching proteins are in the cluster, then
the unbinding time, which again can be equated to the
FRAP recovery time, is simply
τnon−switch ∼ α−1.
(F3)
For typical post-translational modification, or transcrip-
tion termination, this is in the several seconds to minutes
timescale, which is compatible with experimental results.
Aside from FRAP, another important timescale is that
over which local TADs form (e.g., in Fig. 4), τTAD. In
Figure S6: Chromatin contacts for different choices of specific
and non-specific binding affinities. The plot shows the fraction of
non-local versus local contacts for a chromatin fiber; fiber patterning
and all parameters are as in Fig. 4 of the main text. Simulations
initially involved non-switching proteins; half-way through the
simulation, proteins began to switch (α = 0.0001 inverse Brownian
times). Contacts are classified as local (non-local) if they involve
beads separated less than (more than) 400 beads along the chain (or
1.2 M bp. Non-specific (1) and specific (2 interaction energies are
indicated on the right of the plot, in the format (1, 2).
14
analogy with polymer collapse and heteropolymer folding
(see, e.g., Ref. [4]), we expect τTAD, to be a power law in
the number of monomers in the TAD, say M , where the
prefactor should describe microscopic (diffusion) dynam-
ics of a monomer. Dimensional analysis then suggests
τTAD ∼ σ2
D
M z ∼ τBM z
(F4)
where z is a scaling exponent. The Brownian time τB is
of the order of 10−3 s with previous assumptions for vis-
cosity and monomer size, while in our simulations z (cid:39) 1
at least up to M ∼ 100 (corresponding to 300 kbp). Also
for eukaryotic chromosomes, TAD size is between 100 kbp
and 1 Mbp, so M is at most a few hundred. Therefore, if
z = 1, we estimate τTAD to be of the order of 1 s, smaller
than typical modification times -- even assuming a larger
effective value of z (e.g., z = 2 gives at most τTAD of order
of 1 min). Previous large-scale simulations also confirm
that eukaryotic TADs form in minutes [5, 6]. These esti-
mates explain why switching proteins in our simulations
can still form TADs in pretty much the same way as non-
switching proteins, and suggest that the same should also
hold for real chromosomes.
switch at a rate α = 0.0001 inverse Brownian times.
Switching arrests coarsening, and leads to clusters of
self-limiting size in steady state.
Supplementary Movie 3: Parameters for this
Movie are as in Fig. 3 of the main text for the α = 0
case. Chromatin beads are not shown for simplicity.
The movie starts with clusters which have formed
during 104
simulation units following equilibration.
The proteins are colored according to the cluster they
belong to when the movie starts; proteins not in any
clusters at that time are gray. The movie then follows
the dynamics with non-switching proteins, for another
104 simulation units:
it can be seen that colored
clusters persist, therefore photobleaching such a cluster
would lead to little or no recovery of signal in the cluster.
Supplementary Movie 4: As Supplementary Movie
3, but now with switching proteins (α = 0.0001 inverse
Brownian times). Proteins are colored according to the
initial clusters; by the end of the simulations all clusters
have mixed colors. While proteins in clusters recycle,
the cluster retains the same overall shape.
Appendix G: Captions of Supplementary Movies
Supplementary Movie 1: A movie of the simu-
lation shown in Figure 1C of the main text. Proteins
do not switch (α = 0). First a snapshot 104 simulation
units after equilibration is shown: a number of small
clusters have formed. Then the subsequent dynamics
are shown: clusters grow and merge, and coarsening
proceeds indefinitely.
Supplementary Movie 2: A movie of the simu-
lution shown in Figure 1D of the main text. Proteins
Supplementary Movie 5: As in Supplementary
Movie 4, but a zoom on two clusters to show more
clearly clusters retain a "memory" of their shape.
Supplementary Movie 6: A movie of the simulation
shown in Figure 4 in the main text. The first half of the
simulation involves non-switching proteins and lasts 105
simulation units: two clusters form. Proteins are black;
yellow chromatin beads are binding, while blue ones are
non-binding. During the second half, proteins are able
to switch (α = 0.0001 inverse Brownian times); clusters
split and interdomain interactions are suppressed.
[1] Kremer K., Grest G. S. (1990) Dynamics of entangled lin-
ear polymer melts: A molecular-dynamics simulation. J.
Chem. Phys. 92(8):5057.
[2] P.M. Chaikin, P. M., Lubensky, T. C. (2000). Principles
of condensed matter physics, Cambrdige University Press,
Cambridge.
[3] Cross, M. C., Hohenberg, P. C. Pattern formation outside
of equilibrium. Rev. Mod. Phys 65, 851-1112 (1993).
[4] Byrne, A., Kiernan, P., Green, D., Dawson, K. A. (1995).
Kinetics of homopolymer collapse. J. Chem. Phys. 1012,
573-577 (1995).
[5] Michieletto, D., Marenduzzo, D., Wani, A. H. (2016)
folding path-
interphase chromosomes.
Chromosome-wide simulations uncover
way and 3D organization of
arXiv:1604.03041.
[6] Brackley, C. A., Johnson, J., Kelly, S., Cook, P. R., Maren-
duzzo, D. (2016). Simulated binding of transcription fac-
tors to active and inactive regions folds human chromo-
somes into loops, rosettes and topological domains. Nucl.
Acids Res. 44, 3503-3512.
15
|
1002.0663 | 1 | 1002 | 2010-02-03T08:16:19 | Linear dielectric response of clustered living cells | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.CB"
] | The dielectric behavior of a linear cluster of two or more living cells connected by tight junctions is analyzed using a spectral method. The polarizability of this system is obtained as an expansion over the eigenmodes of the linear response operator, showing a clear separation of geometry from electric parameters. The eigenmode with the second largest eigenvalue dominates the expansion as the junction between particles tightens, but only when the applied field is aligned with the cluster axis. This effect explains a distinct low-frequency relaxation observed in the impedance spectrum of a suspension of linear clusters. | physics.bio-ph | physics | Linear dielectric response of clustered living cells
Titus Sandu,1 Daniel Vrinceanu,2 and Eugen Gheorghiu1
1International Center for Biodynamics, Bucharest, Romania∗
2Department of Physics, Texas Southern University, Houston, Texas 77004, USA
(Dated: June 7, 2018)
Abstract
The dielectric behavior of a linear cluster of two or more living cells connected by tight junctions
is analyzed using a spectral method. The polarizability of this system is obtained as an expansion
over the eigenmodes of the linear response operator, showing a clear separation of geometry from
electric parameters. The eigenmode with the second largest eigenvalue dominates the expansion as
the junction between particles tightens, but only when the applied field is aligned with the cluster
axis. This effect explains a distinct low-frequency relaxation observed in the impedance spectrum
of a suspension of linear clusters.
PACS numbers: 41.20.Cv, 87.19.rf, 87.50.C-
0
1
0
2
b
e
F
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
3
6
6
0
.
2
0
0
1
:
v
i
X
r
a
∗Electronic address: [email protected]
1
I.
INTRODUCTION
Particle polarizability governs the electric response for many inhomogeneous systems
ranging from biological cells to plasmonic nanoparticles and depends strongly on both its
dielectric and geometric properties. Analytical models have been reported [1, 2] only for
spherical and ellipsoidal geometries, whereas more complex geometries have been approached
by direct numerical solution of the field equations using, for example, the finite difference
methods [3], the finite element method [4], the boundary element method [5, 6], or the
boundary integral equation (BIE) [7].
In a simplified representation, biological cells can be regarded as homogeneous particles
(cores) covered by thin membranes (shells) of contrasting electric conductivities and permit-
tivities. Complex geometries occur when cells are undergoing division cycles (e.g. budding
yeasts) or are coupled in functional tissues (e.g.
lining epithelia or myocardial syncytia).
In these cases, the dielectric/impedance analysis of cellular systems is far more complicated
than previous models [8 -- 10], which considered suspensions of spherical particles. Intriguing
dielectric spectra [11] reveal distinct dielectric dispersions with time evolutions consistently
related to tissue functioning or alteration, identifying a possible role of cell connectors (gap
junctions) in shaping the overall dielectric response.
A direct relation between the microscopic parameters and experimental data can be
analytically derived only for dilute suspensions of particles of simple shapes, and is rather
challenging for system with more realistic shapes, where only purely numerical solutions have
been available. In this work we demonstrate that a spectral representation of a BIE provides
the analytical structure for the polarizability of particles with a wide range of shapes and
structures. The numerically calculated parameters encode particle's geometry information
and are accessible by experiments.
By using single and double-layer potentials [12], the Laplace equation for the fields inside
and outside the particle is transformed into an integral equation. A spectral representation
for the solution of this equation is obtained providing the eigenvalue problem for the linear
response operator is solved. Although not symmetric, this operator has a real spectrum
bounded by -1/2 and 1/2 [13 -- 15] and its eigenvectors are orthogonal to those of the conjugate
double-layer operator. A matrix representation is obtained by using a finite basis of surface
functions.
2
The true advantage of the spectral method is that the eigenvalues and eigenvectors of the
integral operator provide valuable insight into the dielectric behavior of clusters of biological
cells. The eigenvectors are a measure of surface charge distributions due to a field. Only
eigenvectors with a non-zero dipole moment contribute to the polarizability of the particle.
We call these dipole-active eigenmodes. An effective separation of the geometric and mor-
phologic properties from dielectric properties is therefore achieved [16]. We also show that
for a particle covered by multiple confocal shells, the relaxation spectrum is a sum of Debye
terms with the number of relaxations equal to the number of interfaces times the number
of dipole-active eigenvalues. This is a generalization of a previous result [17] on cells of
arbitrary shape.
Our method is related to another spectral approach which uses an eigenvalue differential
equation [18 -- 21]. This method has been applied to biological problems by Lei et al.
[22]
and by Huang et al. [23]. These authors, however, considered homogeneous cells with much
simpler expression for cell polarizability. The BIE spectral method seeks a solution on the
boundary surface defining the particle, as opposed to the eigenvalue differential equation,
where the solution is defined in the entire space.
In a previous study on double (budding) cells it was shown that before cells separation an
additional dispersion occurs [24]. Moreover, in recent papers [3, 25] numerical experiments
have shown that the dielectric spectra of a suspension of dimer cells connected by tight
junctions exhibit an additional, distinct low-frequency relaxation. Our numerical calculation
shows that the largest dipole-active eigenvalue approaches the value of 1/2 as the junction
become tighter. Although the coupling of this eigenmode with the electric field stimulus
is relatively modest (the coupling weight is about 1-2 %), this eigenmode has a significant
contribution to the polarizability of clusters. Thus the eigenmodes close to 1/2 induce an
additional low-frequency relaxation in the dielectric spectra of clustered biological cells even
though the coupling is quite small. Needle-like objects, such as elongated spheroids or long
cylinders, have similar polarizability features.
In this paper we consider rotationally symmetric linear clusters made of up to 4 identi-
cal particles covered by thin insulating membranes and connected by junctions of variable
tightness. Convenient and flexible representations for the surfaces describing these objects
are provided. The number of relaxations in the dielectric spectrum of the linear clusters,
their time constants and their relative strengths are analyzed in terms of the eigenmodes of
3
the linear response operator specific to the given shape.
II. THEORY
A. Effective permittivity of a suspension
We consider a suspension of identical, randomly oriented particles of arbitrary shape and
dielectric permittivity, ε1, immersed in a dielectric medium of dielectric permittivity, ε0.
The dielectric permittivities are in general complex quantities and the theory described here
applies also for time-dependent fields, providing that the size of a particle is much smaller
than the wavelength. When an applied uniform electric field interacts with the suspension,
the response of the system is linear with the applied field and an effective permittivity for
the whole sample can be measured and is defined by [25 -- 27]:
εsus = ε0 + f
αε0
1 − f α
3
.
(1)
This result is obtained in the limit of low concentration, weak intensity of the stimulus field,
and using an effective medium theory within the dipole approximation. Here f = NV1/V is
the volume fraction of all N particles, each of volume V1, with respect to the total volume
of the suspension V . The averaged normalized polarizability α of a particle is defined as
[25, 27, 28]
α =
1
4π V1 ZV1ZΩN (cid:18)ε1 − ε0
ε0 (cid:19) E (N) · N dΩN dV
(2)
where E (N) is the electric field perturbation created inside the particle under a normalized
applied electric excitation with direction N and dΩN is the solid angle element generated
by that direction. The above normalized polarizability is dimensionless and is obtained by
multiplying the standard polarizability of a particle with the factor 4π/V1. In the following
we will refer only to normalized polarizability, thus, without any confusion, the normalized
polarizability α will be simply called polarizability. The directional average in (2) is equiv-
alent to the averaged sum over three orthogonal axes due to the fact that the problem is
linear with respect to the applied field. The latter is more convenient from computational
point of view.
The electric field inside a particle is obtained by solving the following Laplace equation
4
for the electric potential Φ:
∆Φ (x) = 0,
ε0
Φ+ = Φ− ,
∂Φ
∂n(cid:12)(cid:12)+ = ε1
Φ → −x · N,
∂Φ
∂n(cid:12)(cid:12)− ,
x ∈ ℜ3\Σ
x ∈ Σ
x ∈ Σ
x → ∞
(3)
where ℜ3 is the euclidian 3-dimensional space and Σ is the surface of the particle. The
derivatives are taken with respect to the normal vector n to the surface Σ.
Due to the mismatch between the polarization inside and outside the object, electric
charges accumulate at the interface Σ and create an electric potential which counteracts the
uniform electric field stimulus. The solution of the above Laplace problem (3) is therefore
formally given by
Φ(x) = −x · N +
1
4πZΣ
µ(y)
x − y
dΣ(y)
(4)
The single layer charge distribution µ induced by the normalized electric field is a solution
of the following BIE, obtained by inserting solution (4) in equations (3)
µ(x)
2λ −
1
4πZΣ
µ(y)
n(x) · (x − y)
x − y3
dΣ(y) = n(x) · N
(5)
Here the parameter λ = (ε1−ε0)/(ε1+ε0) isolates all the information regarding the dielectric
properties for this problem.
On using the linear response operator M that acts on the Hilbert space of integrable
functions on the surface Σ,
M[µ] =
1
4π ZΣ
µ(y)
n(x) · (x − y)
x − y3
dΣ(y),
the integral equation (5) is written as
(1/(2λ) − M)µ = n · N.
(6)
(7)
The integral operator (6) is the electric field generated by the single layer charge distribution
µ along the normal to the surface. It encodes the geometric information and has several
interesting properties [13 -- 15]. Its spectrum is discrete and it is not difficult to show that
all of its eigenvalues are bounded by the [-1/2, 1/2] interval. Although non-symmetric, the
operator (6) has real non-degenerate eigenvalues. The eigenvectors are biorthogonal, i. e.,
5
they are not orthogonal among themselves, but orthogonal to the eigenvectors of the adjoint
operator
M †[µ] =
1
4π ZΣ
x − y3
µ(y)
n(y) · (x − y)
dΣ(y),
(8)
which is associated with the electric field generated by a surface distribution of electric
dipoles (double layer charge distribution). Therefore, if uki is a right eigenvector of M
corresponding to eigenvalue χk, Muki = χkuki and hvk′ is a left eigenvector corresponding
to eigenvalue χk′, hvk′M † = χk′hvk′, then
with the scalar product defined as the integral over the interface Σ,
hvk′uki = δk′k,
hf1f2i =ZΣ
f ∗
1 (x) f2 (x) dΣ(x).
(9)
(10)
The value 1/2 is always the largest eigenvalue of the operator M, regardless the geometry
of the object. This is immediately seen if the object is considered to be conductor (ε1 → ∞),
and then the interior electric field has to be zero. In that case λ = 1, and the charge density
that generates a vanishing internal electric field obeys the equation (1/2 − M)µ = 0, and
therefore 1/2 is an eigenvalue of M. However, this eigenmode is not dipole-active and does
not contribute to the total polarization of the object. The operator (6) is insensitive to a
scale transformation, which means that its eigenvalue and eigenvectors depend only on the
shape of the object and not on its size, or electrical properties.
By employing the spectral representation of the resolvent of the operator M
(z − M)−1 =Xk
(z − χk)−1ukihvk,
the solution of equation (7) is obtained for z = 1/(2λ) as
µ =Xk
hvkn · Ni(1/(2λ) − χk)−1uki.
(11)
(12)
The polarizability of the homogeneous particle is obtained by using the distribution (12)
to build the solution (4) of the Laplace equation and use it in equation (2). It has been
shown that, operationally, the polarizability is simply the dipole moment of the distribution
(12) over unit volume [25, 27]
6
α =
1
3
1
V1 Xi,k
hx · Niuki hvkn · Nii
1/(2λ) − χk
,
(13)
where Ni are three mutually orthogonal vectors (directions) of unit norm. The factor
(1/(2λ) − χk)−1 is a generalized Clausius-Mosotti factor. Each dipole-active eigenmode
contributes to α according to its weight pk = 1
V1 Pihx · Niuki hvkn · Nii, which deter-
3
mines the strength of coupling between the uniform electric field and the k-th eigenmode
and contains three components Pk,i = hx · Niuki hvkn · Nii/V1. Equation (13) shows a
clear separation of the electric properties, which are included only in λ, from the geometric
1
properties expressed by χk and pk.
B. Shelled particles
The polarizability of an object covered by a thin shell with permittivity εS can be calcu-
lated in a similar fashion. The electric field is now generated by two single layer distributions,
and boundary conditions are imposed twice, for Σ1 and for Σ2. The surface Σ1 is the outer
surface of the shell and Σ2 is the interface between the particle and the shell. The solution
of a shelled particle in terms of single layer potentials has the form [28]
Φ(x) = −x · N +
1
4π ZΣ1
µ1(y)
x − y
1
4πZΣ2
µ2(y)
x − y
dΣ(y) +
dΣ(y),
(14)
where µ1 and µ2 are the densities defined on surface Σ1 and Σ2, respectively. Four integral
operators M11, M12, M21 and M22 are defined, depending on which surface are variables x
and y. For example M11 is defined when x and y are both on Σ1, M12 is defined by x on
Σ1 and y on Σ2, and so on. Thus
Mij[µj] =
1
4πZΣj
µj(y)
for i, j = 1, 2. The equations obeyed by µ1 and µ2 are
n(x) · (x − y)
dΣ(y)
x − y3
µ1/(2λ1) − M11[µ1] − M12[µ2] = n · N
µ2/(2λ2) − M21[µ1] − M22[µ2] = n · N.
(15)
(16)
Here the electric parameters are: λ1 = (εS − ε0)/(εS + ε0), and λ2 = (ε1 − εS)/(ε1 + εS).
We further assume a confocal geometry, i. e. the surface Σ1 is a slightly scaled version
of Σ2, with a scaling factor η close to unity. This assumption does not provide constant
thickness for the shell, but our main results should remain at least qualitatively valid [3, 6].
7
In the limit of very thin shells, and using the scaling properties of the operator M, one
can show [25, 27] that all four M operators are related to M = M11
M12[µ] = η−3(µ/2 + M[µ])
M12[µ] = −µ/2 + M[µ]
M22[µ] = M[µ].
Equations (16) can then be arranged in a matrix form as
(17)
(18)
1/(2λ1 − M) (1/2 + M)/η3
−1/2 + M 1/(2λ2) − M
µ1
µ2
=
n · N
n · N
.
By knowing the eigenvectors and the eigenvalues of M the charge densities µ1 and µ2 can
be found by inverting the matrix in (18). For example, µ1 is
µ1 =Xk
η3(cid:16) 1
2λ1 − χk(cid:17)(cid:16) 1
η3( 1
2λ2 − χk) +(cid:0) 1
2λ2 − χk(cid:17) +(cid:0) 1
2 + χk(cid:1)
2 + χk(cid:1)(cid:0) 1
2 − χk(cid:1)
hvkn · Niuki.
(19)
The field generated by the the two distributions µ1 and µ2 outside the particle is the
same as the field generated by an equivalent single layer distribution
µe =Xk
hvkn · Ni(1/(2λk) − χk)−1uki,
(20)
where λk = (εk−ε0)/(εk+ε0) and the equivalent permittivity εk is defined for each eigenmode
as:
εk = εS(cid:18)1 +
εS + δ(1/2 − χk)ε1 + δ(1/2 + χk)ǫS(cid:19) ,
ε1 − εS
(21)
where δ = η3−1 ≪ 1. The distribution (20) is similar with the distribution (12) obtained for
a homogeneous particle, except that λ has to be replaced for each mode with an equivalent
quantity λk. Equation (21) can be applied recursively for a multi-shelled structure. The
strict separation of electric and geometric properties is weakened in this case, because the
shape-dependent eigenvalue χk appears now in the electric equivalent quantity λk.
The polarizability of the shelled particle is obtained by using the distribution (20) to
build the solution (4) of the Laplace equation and use it in equation (2), to get
α =
1
3
1
V1 Xi,k
hx · Niuki hvkn · Nii
1/(2λk) − χk
.
(22)
8
Equation (22) is obtained by replacing λ with λk in Eq. (13). The parameter V1 in Eq. (22)
is the total volume of the cell (the core and the shells). In the limit of a dilute suspension
of identical shelled particles, with a low volume fraction f , the effective permittivity (1) is
εsus = ε0 1 + fXk
pk
(1/2 + χk)ε0 + (1/2 − χk)εk! .
εk − ε0
(23)
C. The Debye relaxation expansion
In general, the effective permittivity ǫsus of a suspension of objects with m shells will have
m+1 Debye relaxation terms for each dipole active eigenmode. The proof is recursive and is
based on partial fraction expansion with respect to variable iω of equations (22) and (23),
provided that the complex permittivity of various dielectric phases is ǫ = ε − iσ/(ωεvac)
where i = √−1 and εvac is the permittivity of the free space (8.85 × 10−12 F/m). Thus the
first Debye term comes out from (23) and the remaining m Debye terms result from (21)
by the homogenization process described for shelled particles. Hence, a suspension of cells
with m shells (and m + 1 interfaces) has a dielectric spectrum containing a number of Debye
terms equal to m + 1 times the number of dipole-active eigenvalues.
The suspension effective permittivity ǫsus has the expansion
ǫsus = ǫf +Xk,j
∆εkj/(1 + iωTkj)
(24)
where ǫf = εhf − iσlf/(ωεvac), εhf
frequency conductivity; ∆εkj and Tkj are the dielectric decrement and the relaxation time
is the high-frequency permittivity, and σlf
is the low-
of the kj Debye term, respectively; index k enumerates the dipole-active eigenmodes and
index j enumerates interfaces.
Although the measurable bulk quantities in equation (24) are directly correlated with the
microscopic (electric and shape) parameters, a solution of the inverse problem, which aims
at obtaining the microscopic information non-intrusively, from the effective permittivity, is
in general difficult, if not impossible for the general multi-shell structure. However, biolog-
ical cell has a thin and almost non-conductive membrane, and several simplifications and
approximations can be made. Two Debye relaxation terms in the effective permittivity ǫsus
are expected for each dipole-active eigenmode, corresponding to the two interfaces which
define the membrane.
9
The first relaxation is derived from the equivalent permittivity (21) which can be written
also as Debye relaxation terms:
ǫk = ε + ∆ε/(1 + iωT ).
(25)
The relaxation time T that is given by the poles of ǫk in (21) is a quite good approximation
of the first relaxation time Tk1
T = εvac
(1 + δ/2 + δχk) εS + δ (1/2 − χk) ε1
(1 + δ/2 + δχk) σS + δ (1/2 − χk) σ1 ≈ Tk1.
(26)
The main reason is as follows. At frequencies close to 1/T there is a huge change in ǫk of
order εS/(δ(1/2 − χk)), and consequently a significant change in the total permittivity ǫsus
given by (23). Therefore T provides an approximate value for the relaxation time Tk1 of the
suspension effective permittivity ǫsus.
For a non-conductive shell σS ≈ 0, or more precisely when σS≪δ (1/2 − χk) σ1, the
relaxation time (26) is:
Tk1 ≈ εvac εS/(δ · σ1(1/2 − χk))
(27)
showing a strong dependence on the thickness of the shell and on the shape of the particle,
through the eigenvalue χk. Due to the small parameter δ in (27) the first relaxation (i. e.,
membrane relaxation) tends to have a lower frequency than the second relaxation, which is
present even for particles with no shell (see the discussion below). In addition, cumbersome
but straightforward calculations provide the dielectric decrement ∆εk1 in (24)
∆εk1 ≈ 4f pkεS(δ(1/2 + χk)2(1/2 − χk))−1
(28)
that is very large due to the same strong dependence on the thickness of the shell. The
effect is even more dramatic when the second largest eigenvalue is very close to the largest
eigenvalue, (1/2 − χ2) → 0, like in the case of two cells connected by tight junctions.
For a suspension of shelled spheres η = 1 + ∆R/R and δ = η3 − 1 ≈ 3∆R/R, where
∆R is the thickness of the membrane, R is inner radius, and R + ∆R is the total radius.
Thus both Tk1 and ∆εk1 are proportional to R and εS and inverse proportional to ∆R like
in the Pauly-Schwan theory [1, 8]. Moreover, the dielectric decrement ∆εk1 in (28) is a
generalization of equation (54a) in Ref.
[8].
In the same time, the relaxation time (27)
differs with respect to equation (56a) in Ref.
[8] only by the conductivity term. We will
10
show elsewhere that a more appropriate treatment of the relaxation times recovers also the
relaxation time given by equation (54a) in Ref. [8].
Thus, a non-conductive and thin shell/membrane produces a large relaxation of the com-
plex permittivity of the suspension [31]. The experimental evidences further support these
theoretical facts: when attacking the membrane with a membrane disrupting compound ( for
example a detergent) the relaxation almost vanishes as the cellular membrane is permeated
[32].
For frequencies higher than 1/Tk1 the cell permittivity is essentially determined by the
dielectric properties of the cytoplasm, and does not depend on membrane's properties. The
second Debye relaxation occurs at higher frequencies than the first (membrane) relaxation,
and has the relaxation time
Tk2 ≈ εvac
(1/2 + χk) ε0 + (1/2 − χk) ε1
(1/2 + χk) σ0 + (1/2 − χk) σ1
(29)
derived from the pole of equation (23). The corresponding dielectric decrement is
∆εk2 ≈ f pk(1/2 − χk)(ε1σ0 − ε0σ1)2 ×
((1/2 + χk)ε0 + (1/2 − χk)ε1)−1 ×
((1/2 + χk)σ0 + (1/2 − χk)σ1)−2
The last two equations are similar to the ones that are given for spherical particles in Ref.
[8] (equations (46) and (49) in the aforementioned reference). The relaxation given by Tk2
is basically the relaxation of a homogenous particle embedded in a dielectric environment
and was also discussed in Ref. [22] by a closely related spectral method. If the conductivity
of the cytoplasm is comparable to the conductivity of the outer medium, the decrement of
the second relaxation is small such that it cannot be distinguished in the spectrum. On the
contrary, if the conductivity of the outer medium is much greater or smaller that that of
cytoplasm, than a second observable relaxation occurs. Unlike the membrane relaxation,
this second relaxation depends only weakly on the shape. By assuming that σ0 ≪ σ1
and by using a finite-difference method, this resonance was also obtained in [3] and it was
instrumental in explaining the experimental data on the fission of yeast cells of Asami et al.
[33] by Lei et al. [22].
The shape of the particle is important because it affects the number of dipole-active
eigenvalues and their strengths. In principle, each dipole-active eigenvalue introduces a new
11
relaxation in the dielectric spectrum, providing this relaxation is well separated from the
others. A cluster with complex geometry can have several dipole-active eigenvalues, but
unless the cluster is larger in one dimension then in the others, or there are tight junctions,
the relaxations overlap to create broad features in the spectrum. An extra relaxation is
introduced when the particles are covered by thin membranes. In addition, if (1/2 − χk) → 0
for that eigenvalue, then the shell induced relaxation has low frequency, large relaxation
time and large dielectric decrement. Based on the spectral BIE method, it is therefore
possible to explicitly relate the dielectric spectra of cell suspensions to cell's geometry and
electric parameters, and, even design fitting procedures to evaluate these parameters from
measurements.
III. RESULTS
A. Numerical procedure
The calculation of the effective permittivity for a suspension uses equations (1) or (23),
and reduces then to finding the eigenvalues χk and eigenvectors uki and vki of the linear
response operator M. This problem is solved by employing a finite basis of NB functions
defined on the surface Σ. A natural basis for a surface not far from a sphere is the generalized
hyperspherical harmonics functions
Ylm(x) =
Ylm (θ(x), ϕ(x)) ,
(30)
1
ps(x)
1
ps(x)
where s(x) is related to the surface element through dΣ = s(x) dΩx and dΩx is the solid
angle element.
Another choice could be based on Chebyshev polynomials of the first kind [29]
Tlm(x) =
Tl (θ(x)) eimϕ(x).
(31)
Both bases are complete and orthogonal in the Hilbert space of square integrable functions
defined on Σ.
In this paper, we model the linear cluster of particles as an object with axial symmetry.
We seek to find a surface of revolution for which the thickness of the interparticle joints can
be varied without perturbing the overall shape of the object. We use two representations
12
for the surface Σ: (A) for clusters of two particles we use spherical coordinates {x, y, z} =
{r(θ) sin θ cos φ, r(θ) sin θ sin φ, r(θ) cos θ}, and (B) for clusters with more than two particles
we specify the surface in terms of a function g(z) as {x, y, z} = {g(z) cos φ, g(z) sin φ, z}.
In the case B, the surface element is
and the normal to surface Σ is
dΣ = g(z)p1 + g′2(z) dzdϕ,
p1 + g′2(z)
sin ϕ
−g′(z)
n =
1
cos ϕ
(32)
(33)
.
In the basis of generalized hyperspherical harmonics the operator M has matrix elements
given by
Mlm;l′m′ = δmm′
2π
ZZ0
zmax
ZZzmin
A(z, z′, ϕ−ϕ′)P m
l (cos θ(z))P m′
l′ (cos θ(z′)) eim(ϕ−ϕ′) G(z, z′) dz dz′ dϕ dϕ′
(34)
(35)
(36)
where
and
G(z, z′) =qg(z)g(z′)p(1 + g′(z))(1 + g′(z′))−1,
(g(z) − g(z′)) cos φ − (z − z′)g′(z)
A(z, z′, φ) =
[g2(z) + g2(z′) − 2g(z)g(z′) cos φ + (z − z′)2]3/2
After the angle integration in equation (34) and by using the elliptic integrals given in
the Appendix, the matrix elements are obtained by numerical evaluation of the resulting (z,
z′) double integral using an NQ-point Gauss-Legendre quadrature [29, 30]. Because of the
integrable singularity apparent in the kernel of the operator M in equation (6), the mesh
of z must be shifted from the mesh of z′ by a transformation which insures that there is no
overlap between the two meshes.
The delta symbols δmm′ in equation (34) reflects the fact that we consider only objects
with rotational symmetry in this paper. Moreover, for fields parallel with the cluster axis
m = 0, while m = 1 for perpendicular fields.
The convergence of the results is obtained in two steps. First, the number NQ of quadra-
ture points is increased until the matrix elements of M converge, and then the size NB of
13
the basis set is increased until the relevant eigenvalues χk and their corresponding weights
pk have acquired the desired accuracy. A necessary test for convergence is the fulfillment of
the sum rules Pk Pk,i = 1 and Pi,k χkPk,i = 1/2 with sufficient accuracy [18, 19]. Usually
the convergence is fast in both the number of quadrature points and the size of basis, unless
the system has a tight junction where some care must be considered in order to achieve the
required accuracy of the eigenmodes with the eigenvalues close to 1/2. For a sphere there is
just one dipole-active eigenmode which has eigenvalue χ = 1/6 and weight p = 1, while for
an ellipsoid there is one dipole-active eigenmode along each axis. Fast and accurate solutions
are achieved for spheroids with a basis size of NB =20 and with NQ = 64 quadrature points.
In general, the size of the basis and the number of the quadrature points increase with the
number of cells in the cluster and with the decreasing of the junction size. Thus, for our
numerical examples a basis with NB=35-40 and NQ = 128 quadrature points are enough
for a converged solution in the case of the dimers and NB=50 and NQ = 200 quadrature
points in the case of the clusters with up to four cells.
B. Two cells joined by tight junction
The equation r(θ) = (h+cos2 θ)/(1−a cos2 θ) describes the shape of a two-particle cluster.
Parameter h controls the tightness of the inter-particle junction and parameter a measures
the deviation from a spherical shape. More precisely, h is the radius of the smallest circle
at around the thinnest part of the junction.
Figure 1 shows the effective permittivity for a suspension of particles with the following
parameters: ε1 = 70, σ1 = 0.25 S/m, εS = 6, σS = 0, ε0 = 81, σ0 = 0.374 S/m, volume
fraction f = 0.05, membrane thickness δ = 0.00947275, and a = 0.2. The effective per-
mittivity does not depend on the thickness parameter h when the stimulus electric field is
perpendicular to the cluster axis. However, a new relaxation becomes apparent as h → 0,
for parallel fields [3, 25].
Figure 2 presents the first 7 eigenvalues χk and their weights Pk,1 for a field parallel
to the z axis. As the junction become tighter (h → 0) more eigenvalues become dipole-
active. While all eigenvalues are important in shaping the dielectric spectrum, the second
largest eigenvalue χ2 is crucial to explaining the occurrence of an additional relaxation at
low frequencies, as observed for small h in [3, 25]. Although its weight P2,1 also decreases for
14
500
400
p
r
e
p
300
,
r
a
p
200
100
0
h=0.5
h=0.1
h=0.01
h=0.001
perp
101 102 103 104 105 106 107 108 109 1010
f(Hz)
FIG. 1: (Color online) The spectrum of the effective permittivity of a suspension of dimers with
various junction thickness h, and with parallel and perpendicular field configurations. The sus-
pension permittivity for an electric field perpendicular to cluster axis does not depend on h and is
pointed by an arrow.
small h, this dipole active eigenmode approaches 1/2 as the junction becomes tighter. Thus,
according to equations (27) and (28) the effect of χ2 is "enhanced" due to the presence of a
nonconductive shell (as the case for biological particles analyzed in [3, 25]). Moreover, the
decrease of P2,1 is compensated by the increase of 1/(1/2 − χ2).
The presence of a new relaxation at low frequency along with its relationship with the
size of h has been already singled out in Gheorghiu et al. [25] by using the same method
but without the analysis of dipole-active eigenmodes. Using a finite discrete model [3],
the relaxation was observed before the segregation during cell division, while other papers
[34, 35] fail to relate the size of h to the new relaxation, even though one of them [34]
employs essentially the same method as the one outlined in the present work.
Figure 3 shows the charge distribution associated with the first four eigenvalues for two
distinct values of h. The second eigenmode is an antisymmetric combination of net charge
distributions (monopoles) on each particle of the dimer. The third charge distribution is an
antisymmetric combination of charge distributions with a dipole moment on each part of the
15
h
1st
2nd
3rd
4th
5th
6th
7th
0.6
k
0.4
-
5
0
.
0.2
0.0
0.6
0.4
0.2
0.0
1
,
k
P
10-3
10-2
h
10-1
FIG. 2: (Color online) The largest seven eigenvalues and their weights for a binary cluster with
parameter a = 0.2, as a function of h. The inset shows the shape of the dimer.
dimer and the forth distribution is antisymmetric combination of charge distributions with
a quadrupole moment on each particle. At small h (tight junctions), charge accumulates in
the vicinity of the junction [36].
16
)
(
1
u
1st
)
(
2
u
)
(
3
u
2nd
3rd
)
(
4
u
-1
4th
0
cos( )
1
FIG. 3: The first four eigenvectors for a dimer given by equation r(θ) ; a = 0.2 and h = 0.5 (dotted
line) and h = 0.01 (solid line).
C. Clusters of more than two particles
Smoother yet tight junctions would bring (1/2 − χ2) closer to 0 than sharp and tight junc-
tions. The reason is simple: smoother junctions have the two parts of the dimer farther apart.
17
We have analyzed linear clusters of cells connected by smooth and tight junctions by using
a (z, φ) parameterization which describes a surface by {x = g(z) cos ϕ, y = g(z) sin ϕ, z}.
The construction starts from a dimer shape that resembles the shape of the epithelial cells
like MDCK (Madin-Darby Canine Kidney) cells. An example of such shape, displayed in
figure 4, extends from −zmax to zmax and it can be decomposed in three parts: the left cap
(−zmax ≤ z ≤ −z1), the central part (−z1 ≤ z ≤ z1), and the right cap (z1 ≤ z ≤ zmax).
At position ±z1 the shape function has its maximum. An m-cell linear cluster is ob-
tained by repeating the central part m − 1 times and it extends from −Lm to Lm, where
Lm = zmax + (m − 2)z1. Mathematically, the shape is described by:
for − Lm ≤ z ≤ −Lm + zmax
g(z + (m − 2)z1),
g (mod(z + (1 + (−1)m)z1/2, 2z1) − z1) ,
g(z − (m − 2)z1),
for Lm ≤ z ≤ Lm − zmax,
for − Lm + zmax ≤ z ≤ Lm − zmax
(37)
gm(z) =
where mod(x, y) is the remainder of the division of x by y. For the examples considered
here, the dimer shape function is:
g(z) = 0.01 + 2.32317 z2 − 11.9862 z4 + 40.4045 z6
−74.2226 z8 + 79.142 z10 − 51.8929 z12 + 21.3096 z14
−5.35113 z16 + 0.752147 z18 − 0.045375 z20,
(38)
with zmax = 1.77377 and z1 = zmax/2.
Tables I and II list the most representative dipole-active eigenmodes for a trimer in
perpendicular and parallel fields. Only the parallel field configuration has a dipole-active
eigenvalue close to 1/2, with a relatively small weight.
The results for linear clusters of up to four particles are displayed in Figure 5. The elec-
tric parameters are the same as ones used for dimers in the previous section. An additional,
distinct low-frequency relaxation emerges for clusters with more then one particles, only
when the stimulus field is parallel with the symmetry axis. The relaxation frequency de-
creases, while the intensity of these relaxations increases, as the number of cluster members
increases. This behavior is explained again by the combination of eigenvalues close to 1/2,
with thin non-conductive layers covering the cluster and is consistent with experimental data
on ischemic tissues [11], which reports that the cell separation (closure of gap-junctions) is
responsible for decrease and eventual disappearance of the low-frequency dispersion.
18
TABLE I: Most representative dipole-active eigenmodes and their weights for the trimer in parallel
field.
χk
0.4996
0.40642
0.40448
0.37763
0.26409
0.23809
0.17557
0.13522
0.12165
0.07192
0.06649
0.03479
0.03362
0.0281
0.02377
Pk,1
0.01305
0.07769
0.1391
0.17366
0.20519
0.01839
0.05104
0.05234
0.01066
0.06197
0.01239
0.03899
0.02823
0.04688
0.02519
TABLE II: Most representative dipole-active eigenmodes and their weights for the trimer in per-
pendicular field.
χk
0.1831
0.06964
0.05492
0.0272
0.01674
0.01479
0.00395
-0.01635
19
Pk,2
0.64102
0.01464
0.01343
0.01065
0.05687
0.13179
0.05029
0.03246
)
z
(
g
)
z
(
4
g
-zmax
-z1
0
z1
zmax
-2z1-zmax
-3z1
-2z1
-z1
0
z1
2z1
2z1+zmax
FIG. 4: Smooth construction of a cluster (lower panel) from a dimer (upper panel). The parts
determined by z ∈ [−z1, z1] are "glued" together with the ends of the dimer. The arrows show
where the junctions will be placed in the cluster.
In Figure 6 we plot Pk,1/(1/2 − χk) versus (1/2 − χk), which shows that the number of
dipole-active eigenmodes increases with the number of particles in the cluster. According
to (27) and (28), Figure 6 shows in fact the dielectric decrement versus its corresponding
relaxation frequency for each dipole-active eigenmode of the given clusters. For clusters
of two or three particles, there is one important active eigenmode close to 1/2, while for
clusters of four particles there are two active eigenmodes.
It can be conjectured that for a general linear cluster made of m particles, there are m−1
eigenvalues close to 1/2, of which the largest one is always dipole-active and has the largest
20
2500
2000
p
r
e
p
1500
,
r
a
p
1000
500
0
1 cell
2 cells
3 cells
4 cells
perp
101 102 103 104 105 106 107 108 109 1010
f(Hz)
FIG. 5: (Color online) Effective permittivity for clusters (shown in the inset) of one, two, three,
and four cells connected by tight and smooth junctions. The field is either parallel (solid lines
with symbols) or perpendicular (solid lines only) to the cluster axis. The effective permittivity
either increases strongly with the number of cells for parallel geometry, or does not change for a
perpendicular geometry.
weight.
In fact one can show that for two cells connected by smooth and tight junction
characterized by parameter h, (1/2 − χ2) ∝ h2 when h → 0, or more precisely (1/2 − χ2)
is proportional with the solid angle encompassed by the missing part of a cell when it is
connected with other cell in the dimer. The proof is based on the theorem of the solid angle
[12]. The generalization to a finite cluster is also straightforward to (1/2 − χ2) ∝ h2/m (in
that case the solid angle encompassed by the middle junction is proportional to h2/m. The
weight of the second eigenmode is P2,1 = hx · N1u2ihv2n · N1i/V1.
the surface of the cluster is determined by the function g(z) then, up to a constant factor,
hv2n · N1i ≈ g(0)2 = h2 for two cells connected by smooth and tight junctions. The proof
considers that the second eigenfunction of M † is an antisymmetric combination of constant
If we consider that
distributions on each part of the dimer. This assertion is confirmed in Figure 7. Moreover,
hx·N1u2i/V1 is weakly dependent on h. Therefore, for a parallel setting of the field stimulus,
P2,1/(1/2−χ2), which is the measure of the dielectric decrement of low-frequency relaxation,
21
5 0
4 0
3 0
2 0
1 0
)
-
2
/
1
(
/
1
,
k
P
0
4 c e l
l s
3 c e l
l s
2 c e l
l s
1 c e l
l
10-2
10-1
100
10-4
10-3
1 / 2 -
FIG. 6: (Color online) Pk,1/(1/2− χk) versus (1/2− χk) for clusters of up to four cells. The second
eigenvalue has the largest contribution to intensity of relaxation.
is finite and it increases when the number of cells is increased. The increase of relaxation
decrement when m → ∞ is physically limited by σS≪δ (1/2 − χk) σ1, since the membrane
conductivity is not strictly 0.
Due to cluster's shape and membrane properties, the variation of Pk,1/(1/2 − χk) and
(1/2 − χk) with respect the eigenmode k determines a low frequency relaxation when the
dipole-active eigenvalue χk is close to 1/2. We note here that for dipole-active eigenmodes of
ellipsoids the term (1/2−χk) is called the depolarization factor and has analytical expression
[37]. Prolate spheroids with longitudinal axis much larger than the transverse axis (needles)
have the longitudinal depolarization factor approaching 0 and the transverse depolarization
factor approaching 1/2. More precisely, for a long prolate spheroid, the longitudinal depo-
larization factor scales as (1/2 − χ2) ∝ a2
z, az > ax = ay, as ax → 0. On the other
hand, extensive numerical calculations support the fact that cylinders with the same aspect
x/a2
ratio behave similarly to prolate spheroids [37]. Thus, it is not hard to observe that the
low-frequency relaxation of linear clusters of cells connected by tight junctions is similar
to that of a needle or a thin cylinder as long as the cluster and as thick as the junction.
22
3
0
)
z
(
2
u
,
)
z
(
1
u
-3
0.6
)
z
(
2
0.0
v
,
)
z
(
1
v
-0.6
-1
1
0
z
FIG. 7: The first (dotted line) and the second (solid line) eigenfunction of M (upper panel) and
M † (lower panel) for a dimer whose shape is depicted by dashed line in the lower panel. The
second eigenfunction of M † is an antisymmetric combination of almost constant distributions on
each part of the dimer.
23
On the other hand, the high-frequency relaxation of the cluster shows the relaxation of a
suspension of spheroids with the same volume as the volume of a single cell. Therefore the
dielectric spectrum for a suspension of clusters is the same as the spectrum of a two species
suspension made of thin cylinders and spheroids.
IV. CONCLUSIONS
We present a theoretical framework based on a spectral representation of BIE and able to
calculate the dielectric behavior of linear clusters with a wide range of shapes and dielectric
structures. The theory agrees with the results of Pauly and Schwan for a sphere covered
with a shell [1, 8]. In fact, for spheroids, our theory is the same as the analytical results
of Asami et al.
[2]. We present extensive calculations of clusters with shapes resembling
MDCK cells.
A practical numerical recipe to compute the effective permittivity of linear clusters with
arbitrary number of cells is provided. Examples are given for cluster with shapes described as
r(θ) in spherical coordinates or using (z, ϕ) parameters as {x = g(z) cos ϕ, y = g(z) sin ϕ, z}.
Other studies in the literature used only spherical coordinates representation [25, 27, 35]. A
direct relation between the geometry and dielectric parameters of the cells and their dielectric
behavior described by a Debye representation has been formulated for the first time. Other
work [22], which is based on a closely related spectral method [18 -- 21], found a direct relation
linking the geometry and electric parameters to the dielectric behavior only for homogenous
particles. Moreover, the method used in [22] treats only particles with spheroidal geometry.
We show that the spectral representation provides a straightforward evaluation of the
characteristic time constants and dielectric decrements of the relaxations induced by cell
membrane. We prove that the effective permittivity is sensitive to the shape of the embed-
ded particles, specially when the linear response operator has strong dipole active modes
(with large weights pk). A low-frequency and distinct relaxation occurs when the largest
dipole-active eigenvalue is very close to 1/2. Clusters of living cells connected by tight
junctions or very long cells have such an eigenvalue. Our results also shed a new light on
the understanding of recent numerical calculations [38] performed with a boundary element
method on clustered cells where the low-frequency relaxation is attributed to the tight (gap)
junctions connecting the cells. The method used in [38] does not use the confocal geometry
24
assumption .
The present work has several implications and applications. We emphasize the capabilities
of dielectric spectroscopy to monitor the dynamics of cellular systems, e.g., cells during cell
cycle division, using synchronized yeast cells [3, 11, 39], or monolayers of interconnected
cells [40, 41]. Also the method is able to assess the dielectric behavior of linear aggregates
or rouleaux of erythrocytes, where the ellipsoidal or cylindrical approximations are not
adequate [42, 43].
The proposed representation is a powerful alternative to finite element or other purely
numerical approaches, because it provides the analytical framework to explain and predict
the complex dielectric spectra occurring in bioengineering applications. Extension of this
method to other surfaces of revolution, for example linear clusters with more than 4 particles,
is straightforward providing an adequate parametric equation is available. Finally, in many
cases (e.g. shapes with high symmetry) the method is faster, offers accurate solutions and
last but not least can be integrated in fitting procedures to analyze experimental spectra.
Acknowledgments
This work has been supported by Romanian Project "Ideas" No.120/2007 and FP 7
Nanomagma No.214107/2008.
Appendix: Integration over ϕ
The integrals over (ϕ − ϕ′) are performed with the following elliptic integrals,
Z π
0
1
(a − b cos ϕ)3/2 dϕ =
2
√a − b
1
a + b
E(cid:18)−
2b
a − b(cid:19)
Z π
0
cos ϕ
(a − b cos ϕ)3/2 dϕ =
2
√a − b
1
b (cid:20) a
a + b
E(cid:18)−
2b
a − b(cid:19) − K(cid:18)−
2b
a − b(cid:19)(cid:21) ,
(A.1)
(A.2)
Z π
0
cos2 ϕ
(a − b cos ϕ)3/2 dϕ =
2
√a − b
1
b2 (cid:20) 2a2 − b2
a + b
E(cid:18)−
2b
a − b(cid:19) − 2aK(cid:18)−
2b
a − b(cid:19)(cid:21) ,
(A.3)
25
where K(x) and E(x) are the complete integrals of the first and second kind, respectively
[29].
[1] H. Pauly and H. P. Schwan, Z. Naturforsch. b 14, 125 (1959).
[2] K. Asami, T. Hanai, and N. Koizumi, Japan. J. Appl. Phys. 19, 359 (1980).
[3] K. Asami, J. Phys. D 39, 492 (2006).
[4] E. Fear and M. Stuchly, IEEE Trans. Biomed. Eng. 45, 1259 (1998).
[5] K. Sekine, Y. Watanabe, S. Hara, and K. Asami, Biochim. Biophys. Acta 1721, 130 (2005).
[6] M. Sancho, G. Martinez, and C. Martin, J. Electrostat. 57, 143 (2003).
[7] C. Brosseau and A. Beroual, Prog. Mater. Sci. 48, 373 (2003).
[8] H. P. Schwan and K. R. Foster, in Handbook of Biological Effects of Electromagnetic Fields,
edited by C. Polk and E. Postow (CRC Press, Boca Raton, Florida, 1996), p. 25.
[9] E. Gheorghiu, Phys. Med. Biol. 38, 979 (1993).
[10] E. Gheorghiu, J. Phys. A 27, 3883 (1994).
[11] J. Knapp, W. Gross, M. M. Gebhard, and M. Schaefer, Bioelectrochemistry 67, 67 (2005).
[12] V. Vladimirov, Equations of mathematical physics (MIR, Moscow(Translated from Russian),
1984).
[13] F. Ouyang and M. Isaacson, Philos. Mag. B 60, 481 (1989).
[14] D. R. Fredkin and I. D. Mayergoyz, Phys. Rev. Lett. 91, 253902 (2003).
[15] I. D. Mayergoyz, D. R. Fredkin, and Z. Zhang, Phys. Rev. B 72, 155412 (2005).
[16] D. Vranceanu and E. Gheorgiu, Bioelectrochem. Bioenerg. 40, 167 (1996).
[17] T. Hanai, H. Z. Zhang, K. Sekine, K. Asaka, and K. Asami, Ferroelectrics 86, 191 (1988).
[18] D. J. Bergman, Phys. Rep. 43, 377 (1978).
[19] D. J. Bergman and D. Stroud, Solid State Physics (Academic Press, New York, 1992), vol. 46,
p. 147.
[20] M. I. Stockman, S. V. Faleev, and D. J. Bergman, Phys. Rev. Lett. 87, 167401 (2001).
[21] K. Li, M. I. Stockman, and D. J. Bergman, Phys. Rev. Lett. 91, 227402 (2003).
[22] J. Lei, J. T. K. Wan, K. W. Yu, and H. Sun, Phys. Rev. E 64, 012903 (2001).
[23] J. P. Huang, K. W. Yu, and G. Q. Gu, Phys. Rev. E 65, 021401 (2002).
[24] K. Asami, E. Gheorgiu, and T. Yonezawa, Biochim. Biophys. Acta 1381, 234 (1998).
26
[25] E. Gheorghiu, C. Balut, and M. Gheorghiu, Phys. Med. Biol. 47, 341 (2002).
[26] J. D. Jackson, Classical Electrodynamics (John Wiley Sons, New York, 1975).
[27] C. Prodan and E. Prodan, J. Phys. D 32, 335 (1999).
[28] J. L. Sebastian, S. Munoz, M. Sancho, and G. Alvarez, Phys. Rev. E 78, 051905 (2008).
[29] M. Abramowitz and I. Stegun, eds., Handbook of Mathematical Functions with Formulas,
Graphs, and Mathematical Tables (Dover, New York, 1972), 9th ed.
[30] J. P. Boyd, Chebyshev and Fourier Spectral Methods (Dover, New York, 2001).
[31] H. Fricke, J. Appl. Phys. 24, 644 (1953).
[32] K. Asami, T. Hanai, and N. Koizumi, J. Membr. Biol. 34, 145 (1977).
[33] K. Asami, Biochim. Biophys. Acta 1772, 137 (1999).
[34] A. D. Biasio and C. Cametti, Biolectrochemistry 71, 149 (2007).
[35] A. D. Biasio, L. Ambrosone, and C. Cametti, J. Phys. D 42, 025402 (2009).
[36] V. V. Klimov and D. V. Guzatov, Phys. Rev. B 75, 024303 (2007).
[37] J. Venermo and A. Sihvola, J. Electrostatics 63, 101 (2007).
[38] A. Ron, N. Fishelson, N. Croitoriu, D. Benayahu, and Y. Shacham-Diamand, Biophys. Chem.
140, 39 (2009).
[39] E. Gheorgiu and K. Asami, Bioelectrochem. Bioenerg. 45, 139 (1998).
[40] J. Wegener, C. R. Keese, and I. Giaever, Exp. Cell Res. 259, 158 (2000).
[41] E. Urdapilleta, M. Bellotti, and F. J. Bonetto, Phys. Rev. E 74, 041908 (2006).
[42] J. L. Sebastian, S. M. S. Martin, M. Sancho, J. M. Miranda, and G. Alvarez, Phys. Rev. E
72, 031913 (2005).
[43] K. Asami and K. Sekine, J. Phys. D 40, 2197 (2007).
27
|
1302.0267 | 1 | 1302 | 2013-02-01T20:14:55 | Polar vs. apolar alignment in systems of polar self-propelled particles | [
"physics.bio-ph",
"cond-mat.soft"
] | The symmetry of the alignment mechanism in systems of polar self-propelled particles determines the possible macroscopic large-scale patterns that can emerge. Here we compare polar and apolar alignment. These systems share some common features like giant number fluctuations in the ordered phase and self-segregation in the form of bands near the onset of orientational order. Despite these similarities, there are essential differences like the symmetry of the ordered phase and the stability of the bands. | physics.bio-ph | physics |
Polar vs. apolar alignment in systems of polar
self-propelled particles
Fernando Peruani
Max Planck for the Physics of Complex Systems, Nothnitzer str. 38, Dresden, Germany
E-mail: [email protected]
Francesco Ginelli
Service de Physique de l'Etat Condens´e, CEA-Saclay, 91191 Gif-sur-Yvette, France
Markus Bar
Physikalisch-Technische Bundesanstalt, Abbestrasse 2-12, 10587 Berlin, Germany
Hugues Chat´e
Service de Physique de l'Etat Condens´e, CEA-Saclay, 91191 Gif-sur-Yvette, France
Abstract. The symmetry of the alignment mechanism in systems of polar self-propelled
particles determines the possible macroscopic large-scale patterns that can emerge. Here we
compare polar and apolar alignment. These systems share some common features like giant
number fluctuations in the ordered phase and self-segregation in the form of bands near the
onset of orientational order. Despite these similarities, there are essential differences like the
symmetry of the ordered phase and the stability of the bands.
1. Introduction
Systems of collectively moving entities are ubiquitous in nature, ranging from flocks of
birds [1, 2, 3], insect swarms [5, 6], bacterial collective motion [7, 20], to even driven granular
media [10, 11, 12, 13]. Beyond the complexity of each system, it is possible to classify the
interaction among the moving entities according to its symmetry, which can be either polar or
apolar. Polar (or ferromagnetic) interactions lead to parallel alignment of the velocity of the
moving objects, while apolar (or nematic) interactions allow both, parallel as well as antiparallel
alignment. Fish, birds, and insects seem to exhibit polar alignment [1, 2, 3, 5, 6]. Bacteria and
driven granular media often display apolar alignment [20, 10, 11, 12, 13]. The physical origin of
the alignment mechanism varies among these examples and it is difficult to identify a particular
physical interaction with a given symmetry in these non-equilibrium systems. For instance, steric
interactions among elongated self-propelled objects lead to apolar alignment. This can be easily
observed in realistic self-propelled rod models [15], experiments with driven rods [11, 12], and
experiments with gliding bacteria [20]. However, if the objects are isotropic, volume exclusion
effects can result in a polar alignment, as observed in models [14] and experiments [13] with
polarly driven disks.
Here we focus on the collective large-scale patterns emerging in systems of point particles
moving in two-dimensions at constant speed. We compare the dynamics and macroscopic
properties of self-propelled particles interacting by a polar and an apolar velocity alignment
mechanism. We list the common features and highlight the differences. We provide in this way
a summary of the two main classes of polar self-propelled particle systems.
2. Equation of motion of self-propelled particles
The evolution of the i-th particle is given by the following updating rules:
xt+∆t
i
θt+∆t
i
= xt
i + v0ei θt
i ∆t
= arg
Xxt
i −xt
j, θt
f(cid:0)θt
j
i(cid:1) ei θt
+ ηt
i
(1)
(2)
i is the position of the particle and θt
j≤ǫ
where xt
i its direction of motion at time t, v0 the active
particle speed, arg (b) indicates the argument of the imaginary number b , ηt
i is a delta-correlated
white noise of strength η, and ∆t is the temporal time step. Notice that Eqs. (1) and (2) can
be considered the limiting case of very fast angular relaxation of a system of equations of the
form:
xi = v0ei θi ,
θi = −γ ∂U
∂θi
(xi, θi) + ηi(t) ,
as discussed in [16]. The symmetry of the alignment mechanism is contained in the function f
which is defined as:
j, θt
f(cid:0)θt
i(cid:1) =(1
for polar alignment
sign(cid:16)cos(θt
j − θt
i)(cid:17) for apolar alignment
,
(3)
where sign (x) return a +1 if x ≥ 0 and −1 otherwise. Eqs. (1) and (2) together with Eq. (3)
defined the so-called Vicsek model [17, 18] for polar alignment, and the model for (ideal) self-
propelled rods introduced in [16, 19] for apolar alignment.
2.1. Order parameters
The orientational order can be characterized by the following order parameters. The polar
(ferromagnetic) order parameter is defined by:
exp(i θt
i ,
(4)
where h. . .i denotes time-average and N stands for the total number of particles in the system.
φ takes the value 1 when all particles move in the same direction, while in the disordered phase,
i.e., when particles move in any direction with equal probability, it vanishes. On the other hand,
the apolar (nematic) ordered parameter takes the form:
1
N
N
Xk=1
φ = h(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1
N
N
Xk=1
S = h(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
k)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
k)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
exp(i 2 θt
i .
(5)
Formally, S can be derived from the order parameter matrix Q of liquid crystals (LC) [21], as
the largest eigenvalue (which here we have normalized such that SLC ∈ [0, 1]). When the system
is perfectly nematically ordered, i.e., when particles move in opposite directions along the same
axis, S takes the value 1. In summary, a perfectly polarly ordered phase is characterized by
φ = S = 1, while for a genuine apolarly ordered phase, φ = 0 and S = 1.
(a)
(b)
(c)
I
II
III
Typical steady-state snapshots for SPP with polar alignment at different noise
Figure 1.
values (linear size L = 512, particle density ρ = 1/8, active speed v0 = 1/4). The snapshots
correspond to phase I, (a) η = 0.06, phase II, (b) η = 0.1, and phase III, (c) η = 0.19. The
arrow in (b) indicates the direction of motion of the high-density band.
.
3. Macroscopic patterns with polar alignment
We start out by reviewing the large-scale properties of SPP with polar alignment. If we fix the
active speed v0 and the density ρ, while varying the noise strength η we observe the emergence
of three statistically distinct stationary states or phases. Two of these phases correspond to
orientationally ordered phases, while the third one is a disordered phase. We refer to these
phases as phase I, II, and III by increasing value of η.
Phase I exhibits polar long-range order and is characterized by the absence of regular large-
scale, high density structures, see Fig.1(a). Due to this fact, it is often said that this phase is
spatially homogeneous albeit with large density fluctuations. The presence of polar long-range
order implies that φ(η) → K(η) as the system size L → ∞, with K(η) > 0 a constant that
depends only on η. Phase I also exhibits giant number fluctuations (NF). NF are defined as
∆n2(ℓ) = h(n(ℓ) − hni(ℓ))2i, where n(ℓ) stands for the number of particles in a box of linear
size ℓ. The average number of particles is hni = ρℓ2 while ∆n(ℓ) is expected to be ∆n ∝ hniα.
Giant NF correspond to α > 1
By increasing η we reach a point at which large-scale, elongated, high-density, high-order
solitary structures emerge. These structures, that we refer to as bands, move at roughly constant
speed, see Fig.1(b). A system of SPPs with polar alignment can display multiple bands. The
low-density region in between bands is disordered and there is no characteristic separation length
between bands [18]. In summary, phase II is characterized by polar long-range order and the
presence of travelling bands. If η is increased further, we reach phase III that exhibits local and
global disorder, see Fig.1(c).
2 [22]. In [18] it was shown that α ∼ 0.8 for phase I.
4. Macroscopic patterns with apolar alignment
Now we focus on a system of SPP with apolar alignment. While sweeping η, keeping the active
speed and the density fixed, we observe the emergence of four statistically distinct stationary
states or phases, see Figs. 3-4. Two of these four regimes correspond to ordered phases while
the other two to disordered phases. At the two extremes, very high and very low values of η,
the system is spatially homogeneous, Figs. 3(a,e). At intermediate values of η, spontaneous
density segregation occurs in the form of a high-density ordered region along which the particles
move back and forth, Figs. 3(b-d). It is important to stress that macroscopic polar order always
remains near zero. Nevertheless, if we look at very short length scales, we observe that the
1
0.8
0.6
0.4
0.2
0
0.04
0.03
0.02
0.01
0
φ
φ
.
v
e
D
.
d
t
S
I
II
III
0.1
η
0.2
Figure 2. Polar order parameter φ and its standard deviation as function of the noise amplitude
η (other parameters as in Fig. 1).
(a)
(b)
(c)
(d)
(e)
I
II
III
IV
Figure 3. Typical steady-state snapshots for SPP with apolar alignment at different noise
values (linear size L = 2048, density ρ = 1/8, and velocity v0 = 1/2). (a) η = 0.08, (b) η = 0.10,
(c) η = 0.13, (d) η = 0.168, (e) η = 0.20. Arrows indicate the polar orientation of particles
(except in (d)); only a fraction of the particles are shown for clarity reasons.
apolar ordered phase is formed by polarly oriented clusters. In the following we provide a more
quantitative description of these four phases, which we name, by increasing value of η, phase I to
IV. Phase I exhibits true long-range apolar (nematic) order and is spatially homogeneous. This
apolar order corresponds to two subpopulations of particles of roughly equal size that migrate in
opposite directions (Fig. 5a). We conclude that the apolar order is truly long-range by looking
at scaling of the apolar order parameter S with the system size. From Fig. 5a, it seems that
S(L) approaches a constant asymptotic value C0(η) for L → ∞, with C0(η) > 0, indicating the
possible existence of true long-range apolar order. Another remarkable feature of phase I is the
presence of giant number fluctuations (NF). Fig. 5b shows the scaling ∆n ∝ hniα in phase I,
where it is found again that α ∼ 0.8.
By increasing η we move from phase I to phase II. Its onset is characterized by the emergence
of a narrow, low-density, disordered channel. This channel becomes wider at larger η values,
so that one can speak of a high-density ordered region along which particles travel in both
directions. The high-density region exhibits apolar order with properties similar to those
0.8
0.6
0.4
0.2
S
0
0.08
0.1
S
.
v
e
D
.
d
t
S
0.01
0.001
0.08
I
II
III
IV
0.1
0.12
η
0.14
0.16
0.18
0.1
0.12
0.14
η
0.16
0.18
Figure 4. Apolar (nematic) order parameter S and its standard deviation as function of the
noise amplitude η (other parameters as in Fig. 3).
Figure 5. Phase I of SPPs with apolar alignment: long-range apolar order and giant number
fluctuations (ρ = 1/8 and η = 0.095). (a) Apolar order parameter S vs system size L in square
domains. The vertical red dashed line marks the persistence length χ ≈ 4400 that particles
can travel before performing a U-turn. Inset: S − C0 = 0.813063 vs L (red dashed line: L−2/3
decay).
(b) Number fluctuations ∆n as function of hni. The dashed line correspond to an
algebraic growth with exponent 0.8 (simulations with L = 4096). See Ref. [19].
observed in phase I: true long-range order and giant NF. The (rescaled) region possesses a
well-defined profile with sharper and sharper edges as L increases (Fig. 6a). The fraction area Ω
occupied by the dense region is thus asymptotically independent of system size, and it decreases
continuously as the noise strength η increases (Fig. 6b).
By increasing η even further, we reach phase III, where the dense region becomes unstable
and constantly bend, merge, break and reform. As result of this dynamics, S(t) fluctuates
strongly and on very large time scales, but its average decreases as 1/√N [19]. Thus, phase III
is characterized by the absence of long-range order and the presence of large correlation lengths
and times. At even larger η values, we find phase IV, which exhibits local and global disorder
on smal length- and time-scales, and is spatially homogeneous.
1
S⊥
ρ⊥
0.5
0
0.2
0
0
(a)
(b)
1
Ω
0.5
x/L
η
0.5
0
1
0.1
0.15
Figure 6. Phase II of SPPs with apolar alignment is characterized by the presence of a dense
region. (a) Rescaled transverse profiles in square domains of linear size L = 512 (black), 1024
(red), and 2048 (blue) at η = 0.14. (Data averaged over the longitudinal direction and time,
translated to be centered at the same location.) Bottom: density profiles. Top: nematic order
parameter profiles. (b) Surface fraction Ω as a function of noise amplitude η (defined here as
the width at mid-height of the rescaled S profile). See Ref. [19].
phase IV
η
phase II
phase I
phase III
0.1
0.1
ρ
1
Figure 7. Phase diagram - Critical value of η, separating the four phases, as function of the
density. Thresholds obtained for systems of size L = 1024. The dashed line corresponds to the
scaling η ∝ ρ1/2.
5. Comparison between polar and apolar alignment
Self-propelled particles moving at constant speed with either polar or apolar alignment exhibit
a phase transition from a disordered to orientationally ordered phase. For both alignments the
order seems to be long-range, being purely polar for polar alignment and purely apolar (nematic)
for apolar alignment. It has been shown theoretically that self-propelled particles with polar
alignment can display long-range order [23]. However, the claim about the existence of apolar
long-range order is exclusively based on large-scale simulations and a theoretical justification is
still missing. Assuming that the findings here reported for large-scale simulations hold in the
thermodynamical limit, we conclude that SPPs with polar alignment exhibit three phases, while
with apolar alignment, the phases are four. For both alignments, there are two ordered phases,
one of them spatially "homogeneous" though with giant number fluctuations with roughly the
same critical exponent 0.8, while the other one is characterized by the presence of high-density,
high ordered regions. For polar alignment, the system can display multiple solitary bands.
These bands are travelling bands whose moving direction is perpendicular to the long axis of
the band, i.e., particles are aligned roughly in the same direction exhibited by the band velocity.
On the other hand, for apolar alignment, we observe only one dense region. This region does
not move and is composed of two populations of particles moving in opposite direction along
its periodic axis. For polar alignment there is one disordered phase characterized by local and
global disorder. SPPs with apolar alignment also exhibit a similar phase, however these particles
also display an intringuing disordered phase where particle self-segregation in the form of highly
dynamical bands occur.
This means that in a coarse-grained description of these systems, the homogeneous disordered
phase should get unstable by the emergence of soliton-like structures, the travelling bands, for
polar alignment while for apolar alignment, local alignment has to lead to the formation of highly
dynamical region. In the case of apolar alignment, the onset of global order is not associated to
the instability of the disorder homogeneous state, as for polar alignment, but with the stability
of the dense region. On the other extreme, well in the "homogeneous" ordered phase, we can
approximate the behavior of the order parameter φ and S as:
φ =
S =
sin(η/2)
2
η
1
η2 sin2(η) ,
(6)
(7)
using a simple mean-field argument as discussed in [24]. From this we can presume that for
a given set of parameters v0, ρ, and η, φ(v0, ρ, η) ≥ S(v0, ρ, η). Moreover, a simple mean-field
analysis of the"homogeneous" disordered phase [16] reveals that, for fixed v0 and ρ, this phase
looses its stability at lower values of η for apolar alignment than for polar alignment. From all
this, we learn that is easier to achieve orientational order with polar than with apolar alignment.
This comparison also suggests that polar order is more robust than apolar order. Beyond this
rough comparison between both alignment, a theoretical understanding of the formation of dense
regions in both systems is still lacking. For polar alignment, there are some promising recent
theoretical results [25, 26] that provide an interesting perspective to this puzzling issue. We
hope that in the near future all these open problems will be fully understood.
References
[1] Three Dimensional Animals Groups, edited by J.K. Parrish and W.M Hamner (Cambridge University Press,
Cambridge, England, 1997).
[2] A. Cavagna et al., Proc. Natl. Acad. Sci. 107, 11865 (2010).
[3] K. Bhattacharya and T. Vicsek, New J. Phys. 12, 093019 (2010).
[4] D. Helbing, I. Farkas, and T. Vicsek, Nature (London) 407, 487 (2000).
[5] J. Buhl et al., Science 312, 1402 (2006).
[6] P. Romanczkuk, I.D. Couzin, and L. Schimansky-Geier, Phy. Rev. Lett. 102, 010602 (2009).
[7] H.P. Zhang et al., Proc. Natl. Acad. Sci. 107, 13626 (2010).
[8] V. Schaller et al., Nature 467, 73 (2010).
[9] F. Peruani et al., unpublished.
[10] V. Narayan, S. Ramaswamy, and N. Menon, Science 317, 105 (2007).
[11] A. Kudrolli, G. Lumay, D. Volfson, and L.S. Tsimring, Phys. Rev. Lett. 100, 058001 (2008).
[12] A. Kudrolli, Phys. Rev. Lett. 104, 088001 (2010).
[13] J. Deseigne, O. Dauchot, and H. Chat´e, Phys. Rev. Lett. 105, 098001 (2010).
[14] D. Grossman, I.S. Aranson, and E. Ben Jacob, New J. Phys. 10, 023036 (2008).
[15] F. Peruani, A. Deutsch, and M. Bar, Phys. Rev. E, 74, 030904 (2006).
[16] F. Peruani, A. Deutsch, and M. Bar, Eur. Phys. J. Special Topics 157, 111 (2008).
[17] T. Vicsek, A. Czirok, E. Ben-Jacob, I. Cohen, and O. Shochet, Phys. Rev. Lett. 75, 1226 (1995).
[18] H. Chat´e, F. Ginelli, G. Gr´egoire, and F. Raynaud, Phys. Rev. E 77, 046113 (2008).
[19] F. Ginelli, F. Peruani, M. Bar, and H. Chat´e, Phys. Rev. Lett. 104, 184502 (2010).
[20] Peruani et al., unpublished (2010).
[21] M. Doi and S.F. Edwards, The Theory of Polymer Dynamic (Clarendon Press, Oxford, 1986).
[22] S. Ramaswamy, R.A. Simha and J. Toner, Europhys.Lett. 62, 196 (2003).
[23] J. Toner and Y. Tu, Phys. Rev. E 58, 4828 (1998); Phys. Rev. Lett. 75, 4326 (1995).
[24] F. Peruani, L. Schimansky-Geier, and M. Bar Eur. Phys. J. Special Topics 191, 173-185 (2010).
[25] S. Mishra, A. Baskaran, and M.C. Marchetti, Phys. Rev. E 81, 061916 (2010).
[26] E. Bertin, M. Droz, and G. Gr´egoire, Phys. Rev. E 74, 022101 (2006); J. Phys. A 42, 445001 (2009).
|
1009.0690 | 1 | 1009 | 2010-09-03T14:56:45 | DNA-psoralen: single-molecule experiments and first principles calculations | [
"physics.bio-ph"
] | The authors measure the persistence and contour lengths of DNA-psoralen complexes, as a function of psoralen concentration, for intercalated and crosslinked complexes. In both cases, the persistence length monotonically increases until a certain critical concentration is reached, above which it abruptly decreases and remains approximately constant. The contour length of the complexes exhibits no such discontinuous behavior. By fitting the relative increase of the contour length to the neighbor exclusion model, we obtain the exclusion number and the intrinsic intercalating constant of the psoralen-DNA interaction. Ab initio calculations are employed in order to provide an atomistic picture of these experimental findings. | physics.bio-ph | physics | DNA-psoralen: single-molecule experiments and first principles
calculations
M. S. Rocha1, A. D. L´ucio2, S. S. Alexandre3, R. W. Nunes3, and O. N. Mesquita3
1Departamento de F´ısica, Universidade Federal de Vi¸cosa,
CEP 36570-000, Vi¸cosa, MG, Brazil
2Departamento de Ciencias Exatas, Universidade Federal de Lavras,
Caixa Postal 3037, CEP 37200-000, Lavras, MG, Brazil and
3Departamento de F´ısica, ICEx, Universidade Federal de Minas Gerais,
Caixa Postal 702, CEP 31270-901, Belo Horizonte, MG, Brazil
(Dated: October 30, 2018)
Abstract
The authors measure the persistence and contour lengths of DNA-psoralen complexes, as a
function of psoralen concentration, for intercalated and crosslinked complexes.
In both cases,
the persistence length monotonically increases until a certain critical concentration is reached,
above which it abruptly decreases and remains approximately constant. The contour length of
the complexes exhibits no such discontinuous behavior. By fitting the relative increase of the
contour length to the neighbor exclusion model, we obtain the exclusion number and the intrinsic
intercalating constant of the interaction. Ab initio calculations are employed in order to provide
an atomistic picture of these experimental findings.
PACS numbers: 87.80.Cc, 87.14.gk, 87.15.La
0
1
0
2
p
e
S
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
0
9
6
0
.
9
0
0
1
:
v
i
X
r
a
1
DNA interactions with ligands such as drugs or proteins have been extensively studied
and reviewed by many authors in the past years [1 -- 8]. Some of these drugs, like daunomycin
and ethidium bromide (EtBr), exhibit intercalative binding [1, 3, 6]. Psoralen, by contrast,
presents other forms of linkage to the DNA bases. In fact, it is well known that this drug
can absorb photons and form covalent bonds with the DNA bases, when the sample is
illuminated with ultraviolet-A (UVA) light (λ = 320-400 nm) [4]. Psoralens are compounds
from the family of furocoumarins, broadly used in the medical treatment of various skin
diseases like psoriasis, vitiligo, and some other kinds of dermatitis [9].
FIG. 1: Psol-DNA fragment with five base CG pairs and two intercalated psoralens obtained from
our ab initio DFT calculations.
Although psoralens have been extensively used in medicine for many years, the interac-
tion of the drug with DNA is not yet completely understood. In this work, we study the
intercalative binding and crosslink formation in DNA-psoralen complexes (Psol-DNA), in or-
der to clarify the quantitative aspects of the interaction between psoralen and DNA. We are
interested in probing the modifications of the mechanical properties of DNA molecules when
interacting with psoralen. The elasticity and size changes of these molecules, as a function of
the total psoralen concentration (CP ), are measured by performing single-molecule stretch-
2
ing experiments, which give information about persistence (ξ) and contour (L) lengths of
Psol-DNA. The total psoralen concentration is the sum of the intercalated concentration (Ci)
and the free concentration in solution of psoralen. We also define the intercalated psoralen
fraction r=Ci/Cbp, where Cbp is the DNA base-pair concentration, fixed in all experiments
at 11 µM. In order to elucidate the microscopic mechanisms behind the psolaren-DNA in-
teraction, we also perform first principles calculations for Psol-DNA fragments (an example
is shown in Fig. 1) to determine the behavior of the Psol-DNA stiffness (which is related
to the Young modulus) as a function of CP , and also to obtain the maximum fraction of
intercalated psoralen into DNA, i.e., the limiting value of CP at which a Psol-DNA complex
is still structurally stable.
Experimentally, we observe that as CP increases the Psol-DNA persistence length initially
increases and then undergoes an abrupt transition to a lower value at a critical total psoralen
concentration (C crit
P ), a behavior we also observe in UV-light illuminated complexes.
In
the latter case, the Psol-DNA complex attains higher persistence length values for CP <
C crit
P , which is indicative of crosslinking, where the psoralen binds covalently preferably to
thymine bases in opposite DNA strands, upon illumination. As discussed below, the ab initio
calculations are in qualitative agreement with these experimental findings, enabling us to
analyze the changes in the molecular structure that are responsible for the experimentally
observed behavior.
The intercalated complexes are obtained by simply waiting for the psoralen to interca-
late the λ-DNA in the sample. The crosslinked complexes are obtained by illuminating the
sample with a UVA mercury lamp. We use an optical tweezers to trap a polystyrene bead
attached to the DNA molecule, while pulling the microscope stage with a controlled veloc-
ity. We then make force × extension curves, and use the approximate expression derived by
Marko and Siggia [10] to obtain the persistence and contour lengths of the bare DNA and
Psol-DNA complexes. The persistence length ξ = Y I/kT (Y is the Young modulus, I is the
geometrical moment of inertia of the polymer, and kT is the thermal energy) is a mesoscopic
quantity which depends on the local elasticity of DNA. Its value is ξ ∼ 50 nm for bare DNA
under physiological conditions. In our experiments, the forces applied to stretch the DNA
are within the entropic force regime (forces <∼ 3 pN), thus avoiding externally distorting the
structure and shifting the DNA/psoralen chemical equilibrium (enthalpic effects). The de-
tails about our experimental setup, experimental procedure, and optical-tweezers calibration
3
FIG. 2: Persistence length ξ (in nm) as a function of total psoralen concentration CP (in µM) for
Psol-DNA complexes. C ircles: intercalated Psol-DNA; S quares: crosslinked Psol-DNA. The DNA
base-pair concentration used was Cbp=11 µM.
can be found in Refs. [5, 6, 11].
In an attempt to understand the behavior of ξ as a function of CP , we perform first prin-
ciples calculations of the structural stability and stiffness changes produced by intercalation
of psoralen into fragments of dry poly(dG)-poly(dC) in its acid form. We investigate DNA
models with fragments consisting of four and five guanine-citosine (CG) pairs. A five-base-
pair fragment with two intercalated psoralens is shown in Fig. 1. Our first principles density
functional theory (DFT) simulations use the SIESTA code [15]. Exchange and correlation
effects are described within the generalized gradient approximation [16]. A double-ζ basis set
of atomic orbitals of finite range is used, with polarization functions added for phosphorus
and for the atoms involved in the hydrogen bridges. This methodology has been shown to
provide a good description of the DNA structure [17 -- 19]. We treat isolated finite fragments
in a periodic supercell with large vacuum regions. For the DNA bases at the edges of the
4
fragments, one atom was held fixed in the same position as in the infinite periodic DNA [17].
Broken bonds at the edges are saturated with hydrogen atoms. Full geometry relaxations
were performed (forces < 0.04 eV/A). The stability of Psol-DNA complexes is addressed by
computing their formation energies defined as the difference between the total energy of the
Psol-DNA complex and the sum of the total energies of the bare-DNA fragment and the
isolated psoralen molecule, after full relaxation of the initial geometries.
FIG. 3: Total psoralen concentration CP as a function of relative increase of the contour length
Θ for Psol-DNA intercalated complexes. Θ for psoralen molecules equals the intercalated psoralen
fraction r. Dots are the experimental data and the dashed line is a fit using Eq. 1.
In Fig. 2, we plot the measured values of ξ as a function of CP , for both intercalated
and crosslinked Psol-DNA complexes. The figure shows that ξ increases monotonically
until the critical concentration C crit
P
∼ 13.4 µM is reached. For CP > C crit
P , ξ decays
abruptly to around 50 nm, and remains approximately constant at this value, at least for
the concentration range in our experiments. Note that C crit
P
is the same for both intercalated
and crosslinked complexes. A similar abrupt transition of the persistence length was recently
reported by two of us for other intercalating drugs (daunomycin and ethidium bromide) [6].
5
The fact that Psol-DNA complexes exhibit this same behavior is a strong evidence that this
transition is of general character for intercalating molecules, at least in the low-force regime
(<∼ 3 pN) used in our experiments. This abrupt transition may be caused by local formation
of denaturing bubbles, as proposed in Refs. [12] and [13], which softens the structure of the
Psol-DNA complex. In Ref. [12] large scale molecular dynamics calculations indicate the
formation of denaturing bubbles when pure DNA is stretched above 65 pN forces. Since we
use very small stretching forces it is likely that the abrupt transition we observe is caused
by some intrinsic DNA structural changes induced by the intercalating drug.
In Fig. 3 we show the relative increase of the contour length, Θ = (L − L0)/L0, as a
function of CP for the same intercalated complexes shown in Fig. 2, where L0 is the contour
length for CP = 0. Note that the contour length does not exhibit any abrupt changes.
The dashed line in this figure is a fit using Eq.1 below, derived from the neighbor exclusion
model (NEM) [14] (see Ref.
[6] for details). Considering that each intercalated psoralen
molecule increases the length of DNA by the base-pair distance 0.34 nm [4], Θ is equal to
the intercalated psoralen fraction r. In this case the NEM gives
CP = CbpΘ +
Θ(1 − nΘ + Θ)n−1
Ki(1 − nΘ)n
,
(1)
where n is the exclusion number and Ki is the intrinsic intercalating constant. From the
fitting we determine the parameters n = 1.43 ± 0.13 and Ki = (8.8 ± 2.4) × 104 M−1,
for intercalative binding of psoralen. The large error bars for Θ are due to the fact that
different DNAs were used, and there is a natural length distribution for λ-DNA. The abrupt
transition for ξ occurs around Θ =0.38, while the maximum psoralen intercalated fraction
(1/n) is between 0.64 and 0.77.
Figure 4 shows the behavior of the stiffness, which is proportional to the Young modulus
(for fixed DNA length), as a function of r, computed using the DFT methodology. Note
that the stiffness increases as r increases, and an abrupt decrease is observed for r between
0.40 and 0.50, in reasonably good agreement with the data for the persistence length of
Fig 2, where this transition occurs around Θ = r =0.38. Our calculations indicate that
before the transition two psoralens are well intercalated into the DNA structure, while after
the transition one of the psoralens binds with the citosine, and the hydrogen bond between
the two bases is broken, indicating that local denaturation may indeed be the mechanism
behind the abrupt change in stiffness and consequently in the persistence length. Moreover,
6
FIG. 4: Stiffness as a function of intercalated psoralen fraction (r), from ab initio DFT calculations.
Results for r = 0, 0.25, and 0.50 were obtained using a four-base-pair model. Results for r =
0, 0.20, 0.40, and 0.60 were obtained using a five-base-pair model.
our calculations also indicate that when r increases above 0.60, the intercalated structure
is unstable, which sets an upper bound on the maximum fraction of intercalated psoralen
within this theoretical model, in good agreement with the lower-limit value of 0.64 measured
experimentally. Our interpretation for the microscopic nature of the transition rests on the
good qualitative and quantitative agreement between theory and experiment. We must,
however, not refrain from stating the limitations of our model system, since calculations
in such small fragments are strongly affected by boundary effects. This is reflected in the
large values we obtain for the theoretical stiffness, which are also due to the fact that such
small fragments are much stiffer than the much longer DNA chains in the experiments (The
stiffness × length we compute for bare DNA using the four-base-pair model is about 20%
higher than that of the five-base-pair model.)
In conclusion, the authors observe experimentally that the persistence length of DNA-
psoralen complexes increases with psoralen concentration up to a critical concentration,
7
decreasing abruptly and remaining approximately constant above this concentration, a be-
havior also observed in UV-light illuminated crosslinked complexes. Ab initio calculations
show a similar behavior for the stiffness of psoralen-DNA complexes, and also indicate that
local denaturation and binding between psoralen and a DNA base may be the mechanism
behind this persistence length transition.
The authors acknowledge support from the European Grant EU (FP6-029192), and from
the Brazilian agencies: FAPEMIG, CAPES, CNPq, Instituto do Milenio de Nanotecnolo-
gia/MCT and Instituto do Milenio de ´Optica Nao-linear, Fotonica e Biofotonica/MCT.
[1] H. Fritzsche, H. Triebel, J. B. Chaires, N. Dattagupta, and D. M. Crothers, Biochemistry 21
(17), 3940 (1982).
[2] I. Tessmer, C. G. Baumann, G. M. Skinner, J. E. Molloy, J. G. Hoggett, S. J. B. Tendler, and
S. Allen, J. Mod. Optic. 50 (10), 1627 (2003).
[3] A. Sischka, K. Toensing, R. Eckel, S. D. Wilking, N. Sewald, R. Ros, and D. Anselmetti,
Biophys. J. 88 (1), 404 (2005).
[4] D. W. Ussery, R. W. Hoepfner, and R. R. Sinden, Method. Enzymol. 212, 242 (1992).
[5] M. S. Rocha, N. B. Viana, and O. N. Mesquita, J. Chem. Phys. 121 (19), 9679 (2004).
[6] M. S. Rocha, M. C. Ferreira, and O. N. Mesquita, J. Chem. Phys. 127 (10), 105108 (2007).
[7] G. D. Cimino, H. B. Gamper, S. T. Isaacs, and J. E. Hearst, Annu. Rev. Biochem. 54, 1151
(1985).
[8] M. J. McCauley and M. C. Williams, Biopolymers 91 (4), 265 (2009).
[9] W. McNeely and K. L. Goa, Drugs 56 (4), 667 (1998).
[10] J. F. Marko and E. D. Siggia, Macromolecules 28 (26), 8759 (1995).
[11] N. B. Viana, R. T. S. Freire, and O. N. Mesquita, Phys. Rev. E 65 (4), 041921 (2002).
[12] S. A. Harris, Z. A. Sands, and C. A. Laughton, Biophys. J. 88 (3), 1684 (2005).
[13] M. S. Rocha, Phys. Biol. 6, 036013 (2009).
[14] J. D. McGhee and P. H. von Hippel, J. Mol. Biol. 86 (2), 469 (1974).
[15] J. M. Soler, E. Artacho, J. D. Gale, A. Garcia, J. Junquera, P. Ordejon, and D. Sanchez-Portal,
J. Phys-Condens. Mat. 14, 2745 (2002).
8
[16] J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
[17] P. J. de Pablo, F. Moreno-Herrero, J. Colchero, J. G´omez Herrero, P. Herrero, A. M. Bar´o,
P. Ordej´on, J. M. Soler, and E. Artacho, Phys. Rev. Lett. 85, 4992 (2000).
[18] F. L. Gervasio, P. Carloni, and M. Parrinello, Phys. Rev. Lett. 89, 108102 (2002).
[19] S. S. Alexandre, J. M. Soler, L. Seijo, and F. Zamora, Phys. Rev. B 73, 205112 (2006).
9
|
1712.04835 | 1 | 1712 | 2017-12-13T16:07:20 | How to couple identical ring oscillators to get Quasiperiodicity, extended Chaos, Multistability and the loss of Symmetry | [
"physics.bio-ph"
] | We study the dynamical regimes demonstrated by a pair of identical 3-element ring oscillators (reduced version of synthetic 3-gene genetic Repressilator) coupled using the design of the "quorum sensing (QS)" process natural for interbacterial communications. In this work QS is implemented as an additional network incorporating elements of the ring as both the source and the activation target of the fast diffusion QS signal. This version of indirect nonlinear coupling, in cooperation with the reasonable extension of the parameters which control properties of the isolated oscillators, exhibits the formation of a very rich array of attractors. Using a parameter-space defined by the individual oscillator amplitude and the coupling strength, we found the extended area of parameter-space where the identical oscillators demonstrate quasiperiodicity, which evolves to chaos via the period doubling of either resonant limit cycles or complex antiphase symmetric limit cycles with five winding numbers. The symmetric chaos extends over large parameter areas up to its loss of stability, followed by a system transition to an unexpected mode: an asymmetric limit cycle with a winding number of 1:2. In turn, after long evolution across the parameter-space, this cycle demonstrates a period doubling cascade which restores the symmetry of dynamics by formation of symmetric chaos, which nevertheless preserves the memory of the asymmetric limit cycles in the form of stochastic alternating "polarization" of the time series. All stable attractors coexist with some others, forming remarkable and complex multistability including the coexistence of torus and limit cycles, chaos and regular attractors, symmetric and asymmetric regimes. We traced the paths and bifurcations leading to all areas of chaos, and presented a detailed map of all transformations of the dynamics. | physics.bio-ph | physics | How to couple identical ring oscillators to get Quasiperiodicity, extended Chaos,
Multistability and the loss of Symmetry.
Edward H. Hellena, Evgeny Volkovb*
a Department of Physics and Astronomy, University of North Carolina Greensboro, Greensboro, NC USA
b
Department of Theoretical Physics, Lebedev Physical Institute, Leninsky 53, 119991 Moscow, Russia
* - corresponding author, [email protected]
Declarations of interest: none
Target: Communications in Nonlinear Science and Numerical Simulation
Abstract
We study the dynamical regimes demonstrated by a pair of identical 3-element ring oscillators
(reduced version of synthetic 3-gene genetic Repressilator) coupled using the design of the
'quorum sensing (QS)' process natural for interbacterial communications. In this work QS is
implemented as an additional network incorporating elements of the ring as both the source and
the activation target of the fast diffusion QS signal. This version of indirect nonlinear coupling,
in cooperation with the reasonable extension of the parameters which control properties of the
isolated oscillators, exhibits the formation of a very rich array of attractors. Using a parameter-
space defined by the individual oscillator amplitude and the coupling strength, we found the
extended area of parameter-space where the identical oscillators demonstrate quasiperiodicity,
which evolves to chaos via the period doubling of either resonant limit cycles or complex
antiphase symmetric limit cycles with five winding numbers. The symmetric chaos extends over
large parameter areas up to its loss of stability, followed by a system transition to an unexpected
mode: an asymmetric limit cycle with a winding number of 1:2. In turn, after long evolution
across the parameter-space, this cycle demonstrates a period doubling cascade which restores the
symmetry of dynamics by formation of symmetric chaos, which nevertheless preserves the
memory of the asymmetric limit cycles in the form of stochastic alternating "polarization" of the
time series. All stable attractors coexist with some others, forming remarkable and complex
multistability including the coexistence of torus and limit cycles, chaos and regular attractors,
symmetric and asymmetric regimes. We traced the paths and bifurcations leading to all areas of
chaos, and presented a detailed map of all transformations of the dynamics.
Highlights
2
• Two identical genetic Repressilators are coupled via additional network borrowed from quorum
sensing mechanism of bacterial communications.
• Two-frequency torus, complex anti-phase limit cycles and chaos are found over very large areas of
control parameters.
• Asymmetrical limit cycle coexisting with symmetric chaos is discovered and its evolution is studied.
• The paths to chaos across the regions with multistability are presented.
Keywords:
coupled oscillators, quorum sensing, quasiperiodicity, chaos, multistability, symmetry breaking.
1. Introduction
Coupled oscillators are the most frequently used model to study such collective phenomena as
synchronization [1], wave generation, multistability [2] etc. in all fields of fundamental science
and its applications. Coupling can quench oscillations by either suppressing all oscillators to the
same fixed point (amplitude death) or by creating an inhomogeneous stable steady-state
(oscillation death, see [3] for review) which shares the phase space with other oscillatory
regimes. There are many designs of coupling initiated by studies of real systems: starting from
the old classic observations of pulse-coupled fireflies, direct scalar or vector reagents exchange
between chemical reactors, electric and/or synaptic interactions between neurons, up to very
recent investigations of combined coupling between chemical oscillators [4], half-center
oscillator configurations constituted by two bursters [5], and plant interactions [6], which attempt
to explain biennial rhythm in fruit production. Nonlocal connections between oscillators can lead
to the formation of chimeras (coexistence of coherent and incoherent clusters) in homogeneous
populations, even of very different natures: phase oscillators [7, 8], chemical oscillators [9],
metronome ensemble [10], see [11] for review. However, chimera death is also determined by
the particular coupling mechanism [12].
During the last decade new experimental objects – synthetic genetic oscillators – have attracted
great attention [see e.g. reviews: [13, 14] as a new tool for probing the mechanisms of gene
expression regulation and as a possible instrument for genetic therapy. We explore the ring-type
oscillator which is a well-known circuit in physics and applied technology [e.g. review [15]], and
became very popular in synthetic biology after 2000 under the name of "Repressilator" [16] due
to its actual assembly and insertion into the bacterial cell E.coli. The Repressilator consists of
three genes whose protein products (A, B, C) repress the transcriptions of each other
unidirectionally in a cyclic way (..A--▌B--▌C--▌A..). Recently this circuit has been improved
upon [17, 18], "making it an exceptional precise biological clock" [19]. The electronic analog of
the Repressilator was presented in [20, 21]. The cooperativity of transcription repression, which
is the core process of the Repressilator, is typically described by the Hill function ~α/(1+xn)
3
where x is repressor abundance. The dynamics of coupled Repressilators is very sensitive to the
value of n which controls the steepness of repression and α which determines the amplitude of
the isolated Repressilator.
The effectiveness of a genetic oscillator such as the Repressilator depends on how they can
function collectively, thereby requiring a coupling method. An almost obvious suggestion was to
use the natural bacterial quorum sensing (QS) mechanism, used for cell-cell communication in
bacterial populations [22, 23], as the instrument to synchronize genetic oscillators located in
different cells. The core of QS is the production of small molecules (autoinducer) which can,
first, easily diffuse across the cell membrane and external medium and, second, work to
activate/repress transcription of an intended target gene. By manipulating the positions of the
gene providing the autoinducer production and QS-sensitive promoters controlling the
transcription of other genes in the genetic circuits, one can obtain different coupling types and, as
a result, different sets of collective modes in populations of synthetic genetic oscillators. If the
autoinducer is not only a signal molecule but also works as an integral participant of the
oscillator then its diffusion provides the direct coupling which supports, for example, the wave
propagation in bacterial populations [24].
Alternatively, QS may be implemented in a genetic oscillator as an additional element not
required for the generation of the auto-oscillations. In this case the coupling may be described as
indirect because its activity is mediated by the complex chain: production of autoinducer, its
diffusion and binding with the promoter of the target gene. Such coupling is typically nonlinear
because of bimolecular interaction of autoinducer and target instead of the linear intercellular
diffusion of similar variables in direct coupling. This coupling method has been explored in
model simulations of QS-coupled relaxation oscillators [25, 26] and repressilators [27], and its
ability to synchronize the in-phase regime for detuned populations was demonstrated. Further
work [28] has demonstrated that the model [27] may exhibit more complex collective regimes if
the ranges of key parameters are reasonably extended. Later, in-phase synchronization of
coupled Repressilators in the presence of noise was demonstrated [29]. More complex and
flexible dynamics, which better corresponds to biological diversity, have been demonstrated in
[30, 31].
In this work we couple two Repressilators via the scheme of quorum sensing, as suggested in
[32, 33]. In this version of coupling it is important that the production of signal molecules is
associated with the gene which is located inside the Repressilator ring upstream with respect to
the target gene, which accepts the impact of autoinducer. This means that this coupling should be
classified as "conjugate", because in the ordinary differential equation (ODE) system (1) (see the
next section), the production of the autoinducer and its effect on the target gene are controlled by
the equations for different repressors. Intercellular communication is realized by the simple
diffusion of autoinducer but the coupling as a whole is more complex since it cannot be reduced
to the well-studied diffusion-controlled ODE. Such a coupling seems quite acceptable for real
genetic networks and nevertheless may be constructed in artificial networks of a different nature.
4
Recently [34] we have demonstrated by numerical and electronic simulations that two coupled
identical Repressilators show the development of very flexible dynamics if the n increases up to
n = 4. We found spatially homogeneous and inhomogeneous types of chaos over large areas of
the coupling strength (Q) as well as the set of periodic windows which contains inhomogeneous
limit cycles partially synchronized with i:j winding number, e.g. i=1,2…, j=i+1 despite the
identical nature of the oscillators.
Our main goal here is to trace the routes from self-organized quasiperiodicity to the unusual
collective regimes, including homogeneous chaotic regimes and inhomogeneous limit cycles. We
present the detailed map of isolated and overlapping regimes over a large plane of the key
parameters Q and α for other parameters fixed. It is known that in the model [33] the anti-phase
(AP) limit cycle, which is the single stable attractor under small coupling strength, loses stability
via torus bifurcation if coupling strength increases. The boundaries of stable self-organized
quasiperiodic regimes delineate a very large region in the Q-α plane inside which many dynamic
regimes exist. To present the results in a foreseeable form, we restrict the ratio α-max / α-min to
25, based on a reasonable scatter of the corresponding experimental data on the difference
between weak and strong promoters [35, 36].
Two main scenarios of dynamics inside the (Q-α)-plane after torus creation will be presented.
The first one is realized for the values of α just above α-min, defined to be the smallest α for
which the AP limit cycle loses stable continuity during variation of coupling strength Q. Here,
the evolution of the torus progresses towards chaos but encounters a complex symmetric limit
cycle with 5 return times (LC5:5) in each Repressilator. The torus and the LC5:5 have similar
structures, they co-exist, and after destruction of the torus, the system sits on the limit cycle,
which demonstrates the standard period doubling cascade, leading to chaos if Q and/or α
increase. This symmetrical LC5:5 is an isolated attractor with unclear bifurcation origin and is
unrelated to standard "resonant" cycles on the torus.
The second scenario starts for larger α where the LC5:5 loses stability at smaller Q-values than
does the torus, and torus destruction, rather than period doubling of LC5:5, provides the chaotic
oscillations. Increasing Q in the Q-α plane, chaos meets the stability boundary for the spatially
asymmetric (inhomogeneous) LC1:2, after which both attractors coexist in a wide (Q-α)-band.
Further, at the upper boundary of this band the chaos becomes unstable, converting to the
inhomogeneous LC1:2 which develops alone demonstrating both the period doubling cascade,
ending in spatially homogeneous (symmetric) chaos with very "polarized" time series and torus
bifurcation resulting in inhomogeneous quasiperiodic regimes.
Both scenarios are investigated in detail by numerical bifurcation analysis as well as by direct
calculations of time series and Lyapunov exponents. Although we used identical and fairly
smooth oscillators, the proposed QS-coupling scheme generated an unusual cascade of complex
homogeneous and inhomogeneous solutions coexisting in many areas of the phase diagram.
2. Model and Methods
5
We investigate the dynamics of two Repressilators interacting via repressive QS coupling as
used previously [32, 33]. Figure 1 shows two repressilators located in different cells and coupled
via QS to the external medium. The three genes in the loop produce mRNAs (a, b, c) and
proteins (A, B, C), and they impose Hill function inhibition on each other in cyclic order by the
preceding gene. The QS feedback is maintained by the AI produced (rate kS1) by the protein B
while the autoinducer (AI) communicates with the external environment and activates (rate κ in
combination with Michaelis function) production of mRNA for protein C, which, in turn, reduces
the concentration of protein A resulting in activation of protein B production. In this way the
protein B plays a dual role of direct inhibition of protein C synthesis and AI-dependent activation
of protein C synthesis, resulting in complex dynamics of the repressilator, even for just a single
repressilator [37].
Fig.1. A genetic network from two repressilators with QS feedback and the global coupling due to the mixing of
signal molecules in extracellular medium. Lower case (a, b, c) are mRNAs and upper case (A, B, C) are expressed
protein repressors. Si is the autoinducer molecule which diffuses through the cell membrane.
The original models of a single repressilator [16, 33] used re-scaled dimensionless quantities for
rate constants and concentrations. We reduce the model for the case of fast mRNA kinetics ((a,
b, c) are assumed in steady state with their respective inhibitors (C, A, B), so that da/dt = db/dt =
dc/dt ≈ 0). The resulting equations for the protein concentrations and AI concentration S are,
(1a)
(1b)
(1c)
(1d)
where i = 1,2 for the two repressilators, βj (j = 1,2,3) are the ratios of protein decay rate to
mRNA decay rate, α accounts for the maximum transcription rate in the absence of an inhibitor,
niiiCAdtdA11niiiABdtdB12iiniiiSSBCdtdC113extiiSiSiSSBkSkdtdS10
6
and n is the Hill cooperativity coefficient for inhibition. For the quorum sensing pathway kS0 is
the ratio of the AI decay rate to mRNA decay rate, and as previously mentioned, kS1 is the rate of
production of AI and κ gives the strength of AI activation of protein C. The diffusion coefficient
η depends on the permeability of the membrane to the AI molecule. The concentration of AI in
the external medium is Sext and is determined according to quasi-steady-state approximation by
AI produced by both repressilators (S1 and S2), and a dilution factor Q.
(2)
Numerical simulations are performed with XPPAUT [38], AUTO-07p [39], and by direct
integration with 4th-order Runge-Kutta solver. We choose parameter values similar to ones used
previously. Here we use β1 = 0.5, β2 = β3 =0.1, n = 3.0, k=15, kS0 = 1, kS1 = 0.01, and η = 2.
Parameters α and Q are chosen for bifurcation analysis as they represent the amplitude of the
individual oscillators, and the strength of the coupling.
3. Results
3.1 Low-strength Oscillators: α < 1000
To start we present the basic coarse-grained map of regular oscillating regimes and steady states
in system (1), which may coexist with new regimes discussed below or be located in the
parameter space around them. One-parameter XPPAUT continuations shown in Fig. 2 reveal
three branches; a homogeneous steady state (HSS) spanning the Q-range, an inhomogeneous SS
(IHSS) arising from pitchfork bifurcations of HSS (BP1 and BP2), and an anti-phase LC (APLC)
arising from the Andronov-Hopf bifurcation (HB). The steady-states have stable (red) and
unstable (thin black) sections, with the stable IHSS occurring between limit points LP2 and LP3.
There is a narrow range between HB and LP4 of stable low level HSS which coexists with the
stable high value HSS. The APLC is stable (green) at low and high Q-values, and is unstable
(thin blue) between the torus bifurcations of APLC (TR1 and TR2). The APLC is the sole stable
attractor at low Q (see [32] for details).
221SSQSext
7
FIG. 2. The amplitudes of APLC (green line) and HSS, IHSS steady states (red lines) as function of Q. Solid (thin)
lines are stable (unstable) solutions. TR, LP, BP, HB are designations of torus, limit point, branch point (pitchfork),
Andronov-Hopf bifurcations, respectively. n=3, k=15, α=800. Other parameters are fixed: β1 = 0.5, β2 = β3 =0.1, n
= 3.0, kS0 = 1, kS1 = 0.01, and η = 2
The main goal of the current work is the study of dynamics inside the region with unstable
APLC. Figure 2 shows that this area is large for the given parameters. Its size depends on the
values of other model parameters but it will be large throughout our work and the structure of the
phase diagram in Fig.2 will not be changed qualitatively. Any new attractors share the phase
space with HSS and IHSS but the basins of steady states are not so large as to suppress the
exhibition of new regimes.
We choose n=3, k=15 to study the evolution of dynamics starting from torus creation for
increasing coupling strength. Not far from the TR bifurcation the torus is smooth and
demonstrates anti-phase type of oscillations as seen in Fig. 3a.
Fig. 3. Phase portraits of coexisting torus (a) and LC5:5 (b), n=3, α=400, Q=0.64
8
The boundary of the torus stability in the (Q-α)-plane has been calculated by direct integration
and/or by the calculations of Lyapunov exponents which clearly capture the torus death. The fate
of the system (1) after torus destruction depends on the value of α and is presented below.
The unexpected event in the system dynamics is the appearance of the symmetric limit cycle
with 5 winding (rotation) numbers, LC5:5 (Fig. 3b), which is not a classical resonance on the
torus. It coexists with the torus, and its dynamical behavior as a function of Q is controlled by
the value of α. Figure 4 shows the family of Q-continuation plots of the LC5:5 for different
values of α. Also shown is the resonant LC5:5-R which appears as the low-Q pieces for α = 400
and 800. For α = 400 the LC5:5-R exists for only a very narrow Q-range, appearing as a small
spot in Fig. 4. Here we focus our interest on the LC5:5 and its development as α increases.
Fig. 4. The amplitude of oscillations for LC5:5 as a function of Q for α =220, 400, 800. Also shown are resonant
LC5:5-R occurring for α =400 (near the start of Arnold's tongue, see Fig.5) and 800 (two parameter continuations of
both LPs, see in Fig.5). BP is the branch point (pitchfork bifurcation) leading to the loss of symmetry of LC5:5; PDs
mark the start of period doubling cascades of asymmetric LC5:5.
For the small α=220, the LC5:5 is continuously stable and coexists with the torus over the extent
of the LC5:5. Increasing α dramatically changes the dynamics of the LC5:5, which, first, loses
symmetry in pitchfork bifurcation (BP in Fig.4) and, second, goes to chaos via period doubling
cascade as seen for α=400, 800 in Fig. 4. At the large Q>1 the system returns to stable APLC
using the reverse sequence of bifurcations. At the large α=800, the appearance of additional PD
bifurcations within the chaos produces a narrow region of stable LC5:5, which is not studied in
this work. To reveal the dynamics in detail and over broad intervals of key parameters, 2-
parametric (α, Q)-continuations were calculated for the boundaries of torus stability, the
9
bifurcations (LP, BP, PD) of the basic regime LC5:5, and other selected influential limits cycles.
Figure 5 shows the resulting map of regimes.
Fig. 5. The phase diagram demonstrating the boundaries of stability for basic regimes. For 3:4LC and 3:5LC which
are located in intrachaotic periodic windows (PW), separation between LP and PD lines is not apparent at this
resolution.
The order of bifurcations looks like a "matryoshka" in the (Q-α)-plane, with nested torus, LC5:5,
LC3:4, LC2:3, and LC3:5, with the oddity of the strip of overlap due to coexistence of torus and
LC5:5. Also shown is the resonance LC5:5-R, a classical Arnold tongue, which is represented
here to distinguish the positions and the roles of the two L5:5s. Chaos appears on the (Q-α)-map
for α > 350 due to the PD cascades of the nonresonant LC5:5. The main impression from this
map of regimes for our system of two identical coupled oscillators is the vast area between these
PD cascades occupied by chaos. Consider the routes to this chaos.
For 350 < α < 635 (according to Fig. 5), torus destruction takes place before period doubling of
the LC5:5. After its destruction, the system goes into the LC5:5, the single attractor creating
chaos during further Q growth (see Fig. 6 for α=600).
10
Fig.6. Time series B1, B2 of the torus destruction and the system transition to slightly asymmetric LC5:5, n=3, α
=600, Q=0.5685.
The LC5:5 loses symmetry via pitchfork bifurcation, then undergoes a PD-cascade to chaos. For
α = 600, the torus "converts" directly to the As-LC5:5 because the end of torus stability occurs at
slightly higher Q than does the pitchfork bifurcation (BP in Figs. 4, 5) which creates the As-
LC5:5 from the symmetric one. (For smaller α, the torus stability ends before the BP, meaning
the torus "converts" to the symmetric LC5:5). The evolution of chaos is presented as the series
of sequential period maps with increasing complexity (Fig. 7):
Fig. 7. The evolution of sequential period maps T(n+1) vs T(n) after the period doublings of AsLC5:5 up to the
mature chaos for α=600. The value Q=0.58 is the end of LC5:5 PD cascade, (a) Q=0.592, (b) Q=0.604 (c) Q=0.614,
(d) Q=0.634, (e) Q=0.654, (f) Q=0.744. T(n) are the set of Poincaré return times calculated as the time intervals
between the intersections of ascending trajectories of B1(t), B2(t) with the line B = 7.
11
For α > 650, the end of torus stability is beyond the PD-cascade thereby allowing both the torus
destruction and the period doubling of LC5:5 to serve as routes to chaos. Although the Q-interval
of these attractors' coexistence is small, it is interesting to trace the development of chaos in the
presence of the torus covered by resonant cycles. For example, Figure 8 shows the behavior of
two sets of Lyapunov exponents for α = 800: black, blue, pink solid lines correspond to the
calculations started from LC5:5 while red, green, navy dashed lines demonstrate LEs of regime
started from the torus.
Fig.8. Two sets of three LEs as a function of Q: start from 5:5LC (solid black, blue, pink lines), start from torus
(dashed red, green, navy lines). Some winding numbers for resonant cycles on torus and the cycles in periodic
windows in chaos are presented. ETR marks the end of torus. α = 800, Q-step=0.0001. Here and in other Figures for
LEs for two regimes the values for one attractor were artificially shifted upward by 0.0001 or 0.0002 to distinguish
lines near LE=0.
The plots of the two main LEs (solid black and dashed red lines in Fig. 8) show that the LC5:5
and the torus are stable and coexist up to the chaotization of the LC5:5 near Q=0.5015 (positive
bump of black curve). This weak chaos we denote as 5:5-chaos. The torus remains stable as Q
increases, however, the Q-region close to the torus destruction (marked by ETR in Fig. 8)
contains many resonant cycles, which manifest themselves as several dips of the dashed green
curve (LE2-TR). The winding numbers of two resonant cycles are marked in Fig.8, with the
LC23:23 being the starting point of torus destruction. Thus, there is interesting coexistence of the
torus and the 5:5-chaos.
To illustrate the torus transition to chaos we investigated sequential period maps T(n+1) vs. T(n)
in the region of the torus destruction. The first map for Q before transition (Fig. 9a) is not
smooth but it is, nevertheless, a closed curve while the second map (Fig. 9b) contains additional
12
points. Their positions in Fig. 9b indicate the start of torus destruction, after which the evolution
of chaos is very similar to that presented above for α=600 (Fig.7b).
Fig. 9. T(n+1) vs. (Tn) maps: (a)- for torus at Q=0.5155 just after LC23:23; (b)- for transient trajectory between
torus and chaos Q=0.5172. α = 800.
The coexistence of chaos and a non-smooth torus bearing many resonance cycles seems to be a
non-trivial dynamical phenomenon. The Q-interval between the chaotization of LC5:5 and the
end of the stable torus becomes longer if parameter α increases (see lines 5:5PD and TR-end in
Fig. 5) thereby opening other routes to chaos.
Increasing α to 900, we find that the 5:5-chaos is no longer stable over the entire interval to the
ETR. The coexistence ends at the death of the 5:5-chaos, where the system transitions to the
torus. Figure 10 shows the time series of the transition from weak 5:5-chaos to torus, and its
sequential period map. The map contains points for both regimes: the closed curve with small
loop is the map of torus while the combination of the five separate elements is the map of the
5:5-chaos just before the loss of its stability. The torus is then stable over a small Q-interval until
the appearance of a resonant LC23:23 (Q=0.4925) followed by the system's chaotization similar
to that presented in Fig. 9b for α=800.
Fig. 10. The system transition from 5:5-chaos to torus for α=900, Q=0.4808: (a)-time series; (b)- sequential period
map T(n+1) vs (Tn), closed curve – torus map; five separate pig-tails – map of the weak 5:5-chaos.
13
Increasing α beyond 900 causes the evolution described above for α=900 to be interrupted by the
appearance of the relatively broad (Q-α)-island created by the asymmetrical cycle As9:9LC seen
on the top left in Fig. 5 (green line) and shown zoomed-in in Fig. 11.
Fig. 11. The zoom of the (Q-α)-map where the route to chaos via AsLC9:9 takes place.
As an overview of the effect of the AsLC9:9 island we consider α=1000. Figure 11 indicates that
as Q increases from the 5:5PD, the LC5:5 undergoes period-doubling to chaos which coexists
with the As9:9LC until the chaos becomes unstable and the system then transitions to As9:9LC
at Q about 0.458. Prior to the transition, the degree of chaos formation may be estimated from
the sequential period map (Fig. 12a) while the system transition from the chaos to the As-LC9:9
is visualized in Fig. 12b.
Fig. 12. (a) T(n+1)vs T(n) for Q=0.4577 just before the weak chaos transition to asymmetric cycle LC9:9; (b) the
time series of this transition after very small increase of Q. α=1000.
The Lyapunov exponents in Fig. 13 for trajectories started from the LC5:5 and from the torus
show the short Q-interval of coexistence of the weak chaos originated from LC5:5 and the As-
9:9LC, followed by the death of the chaos at Q=0.458 leaving the As-9:9LC as the sole stable
dynamical behavior. The As9:9LC then loses stability near Q=0.478 generating chaos which
gradually develops into the mature chaos with sequential period map like that presented in the
Fig. 7.
14
Comparing the LE graphs for α=800 and 1000, Figs. 8 and 13, it is seen that the As9:9LC at
α=1000 plays the same role as the torus at α=800 regarding coexistence with 5:5-chaos, and
being the sole attractor during final evolution to chaotization. The similar roles are expected
since the As9:9LC is an unusually broad island inside the torus.
Fig. 13. Two sets of LEs for trajectories, started from LC5:5 (solid black, blue, pink lines) and from torus (dashed
red, green, navy lines), as a function of Q, α=1000.
To investigate more thoroughly the route to chaos after crossing the regime of stable AsLC9:9,
the LEs for torus continuation for α=930 have been calculated with a very small Q-step
(0.00001) and are shown in Fig. 14.
Fig.14. The detailed evolution (Q-step is 0.00001) of LEs starting from AsLC9:9 up to the appearance of
irreversible chaos formation after the period doubling of LC23:23, α=930.
15
Behavior of the main LEs shows that the AsLC9:9 is stable up to its LP (Q=0.4811) after which
two LEs are zero demonstrating the system transition to the ergodic torus. However, the torus is
not clearly exhibited because it is covered by tightly packed resonant cycles whose beginnings of
chaotization are unsuccessful until the period doubling of the LC23:23 which leads to creation of
the robust chaos after Q=0.492 where the first LE amplitude surpasses 0.002. This scenario of
the torus with tightly packed resonant cycles leading to chaotization at the ETR occurs for α up
to about 1000. It is different from the LC5:5 period doubling chaotization which dominates for α
< 800. However, after evolution of the chaos with increasing Q, both ways produce similar
chaotic regimes according to the values of LEs and the structure of recurrence time maps.
After chaotization the appearance of many periodic windows is observed over the vast (Q-α)-
area. In most of them limit cycles are symmetric with identical winding number, LCn:n and
identical amplitudes of oscillations in the two oscillators. However, the dynamics in some
windows deserve special attention because they contain spatially asymmetric limit cycles which
are not typical for identical oscillators. The boundaries of AsLC3:4, AsLC2:3, and AsLC3:5 are
pictured in Fig. 5 while the periodic windows with other partial asymmetric LCs are too narrow
to be found with the chosen numerical accuracy used in Fig. 5. Although the bands with stable
AsLCs are narrow (see, for example, interval between 2:3LP and 2:3PD in Fig. 5) their unstable
branches cover large areas of the (Q-α)-plane forming asymmetrical elements in the chaotic time
series. A typical example of the chaotic trajectories with fluctuations of asymmetric "polarity" is
presented in Fig. 15(a) as well as a zoom in Fig. 15(b) showing the long lived unstable orbit
LC2:3. It is notable that for two coupled identical oscillators, the version of QS coupling used in
system (1) has disrupted the homogeneity of oscillations inside chaos and has generated very
different values of recurrence times.
Fig. 15. Time series of symmetric chaotic regime for n=3, α =1000, Q=0.58 (a) and a zoom of "polarized" part from
this time series (b) showing the long-lived unstable 2:3LC.
Thus, for α < 1000 two identical Repressilators indirectly coupled via diffusion of autoinducer
demonstrate several important dynamical features: (i) the existence of a very large area of
parameter-space where the appearance of self-organized quasiperiodicity occurs, which is
unusual for coupled identical oscillators; (ii) the coexistence of torus and non-resonant LC5:5,
each of which initiate the development of chaos in different large areas of the parameter plane;
(iii) the existence of periodic windows with asymmetric limit cycles, the unstable branches of
16
which cover large regions of the parameter space, ensuring the "polarization" of chaotic time
series.
Further growth of α (next section) stimulates the appearance of the other asymmetric limit cycle
1:2, which, in turn, produces the coexistence of regular, quasiperiodic and chaotic, symmetric
and asymmetric regimes over very large parts of (Q-α)-plane.
3.2 High-strength Oscillators: α > 1000
In the previous section, we saw that the main LC5:5 and the torus provide routes to chaos for
most of the region for α<1000. For 800<α<1000 the weak chaos that arises from the period
doubling of the LC5:5 is stable for only a short Q-interval (see Fig. 13). When this chaos
becomes unstable the system transitions to either the torus or As9:9LC depending on the location
in the (Q-α)-map. However, the As9:9LC becomes unstable with further increase of Q and the
system transitions to the torus which then provides the final stage of the route to chaos emerging
at ETR.
Figure 16 shows the region of the (Q-α)-map in which the LC5:5-Res and the As9:9LC island
are the important dynamics interacting with the torus to provide routes to chaos as α increases
beyond 1000. Surprisingly, the external boundary of AsLC9:9 has a smooth closed form and is
limited to α<3200, while the internal marking of stability areas is delineated by additional LP
and PD lines.
Fig. 16. (Q-α)-map of the AsLC9:9 islands, 1:2LP- the boundary of LC1:2 stability, and the resonance LC5:5-R:
(a) boundaries of basic regimes and (b) zoom showing the structure of torus and AsLC9:9 locations (compare with
Fig.17).
The route to chaos for α=1200 is shown by the LEs graph in Fig. 17, in agreement with the
boundaries encountered along the α = 1200 line in Fig. 16b. The LC5:5-Res loses stability at its
LP and the system transitions to the torus as indicated by a second LE with zero amplitude. The
torus has the set of resonant limit cycles (e. g. 14:14, 23:23) and the asymmetric LC9:9 at the
negative dips of the second (green) LE. In contrast to the case in the previous section for α
=1000 where LC9:9 is continuously stable, for α = 1200 there are 3 narrow stable LC9:9 regions
separated by torus and by chaos (main LE > 0). Figure 16b shows the stable torus region
17
delineated by interior 9:9LP structure separating the first two stable LC9:9 regions (see the small
oval in Fig. 16b), and the period-doubled chaos region separating the next two. The chaos
emerging from the period-doubling cascade of the LC9:9 matures, as indicated by the growing
main LE. The growth is interrupted by the narrow periodic window which contains the period-
halving cascade back to stable 9:9LC in the Q-span just prior to the high-Q LP.
Fig. 17. LEs as a function of Q (step=0.0001) for trajectory started from resonant LC5:5-Res. α=1200.
As α increases up to 2200, the Q-interval with stable AsLC9:9 shortens (Fig. 16) and the chaos
emerged from period-doubling starts at smaller coupling strengths. This symmetric chaos
extends to larger coupling strengths where it encounters and coexists over a broad Q-span with a
new asymmetric limit cycle, AsLC1:2, shown in Fig. 16 with dark green line. The time series of
both regimes are presented in Fig. 18 which demonstrates the sharp transition from symmetric
chaos to the AsLC1:2 which occurs at the end of the coexistence region.
18
Fig. 18. The time series of transition from chaotic regime to stable LC1:2 at Q=0.568, for α=2000. The small
fluctuations of amplitudes after the transition are the transient.
By analogy with Fig. 4 for the LC5:5, we calculated for different α, the family of LC1:2 Q-
continuations shown in Fig. 19 which delineates the areas of stability and reveals the basic
bifurcations of this new limit cycle.
Fig. 19. The lower branch of one-parameter Q-continuations of LC1:2 amplitude for α =1800 (LC1:2 is stable from
LP to TR) and for α=3000 with TR bifurcations and PD cascades.
Throughout a large region of coupling strengths the external boundaries of 1:2LC stability are
determined by a limit point at the low-Q end and torus bifurcation at the high-Q end, while the
internal structure of the phase diagram is mainly controlled by the locations of torus and period
doubling bifurcations.
Two-parameter continuations of bifurcation boundaries of important limit cycles, including the
LC1:2 in Fig. 19, are presented in Fig. 20, which is an extension of Fig. 5 for large α. The
boundary of stability of the symmetrical chaos, which coexists with LC1:2, is calculated by
direct integration of the time series or LEs and is presented in Fig. 20 as a black dashed line.
It is noteworthy that over a wide range of α, from 1200 to 2500, chaos is not restored in the
central range of Q-values (0.65 to 0.95). Instead, after the end of the chaos emerged from PD of
AsLC9:9, there is a broad area dominated by stable LC1:2 and LC2:4. It is only when α
surpasses 2500 when the PD-cascade of the LC1:2 creates chaos.
19
Fig. 20. Q-α regime map for α > 1000. Only those boundaries relevant to chaos evolution are shown. The white
area adjacent to LC1:2 and LC2:4 contains many asymmetric limit cycles, see Fig. 22 as example.
For α > 2500 and the intermediate values of coupling strength (0.65<Q< 0.85), the system (1)
demonstrates the classic route to chaos via period doubling cascades of the AsLC1:2. The
chaotization after period doublings results in the restoration of spatial symmetry of the attractor
as a whole because the amplitudes of the Repressilator's oscillations become the same.
However, this formal symmetrization masks the real temporal asymmetry of the chaotic
trajectories. The typical trajectory presented in Fig. 21 clearly demonstrates the stochastic long-
lasting intervals with the fixed polarity.
20
Fig. 21. The random example of the time series demonstrating the chaotic switching of polarity, α=3000, Q=0.797.
The existence of long almost deterministic asymmetric orbits within a chaotic time series can
change the widely-held view of chaos as a short-correlated irregular process. Because the
asymmetric LC is the starting attractor for the development of chaos, there are large parameter
regions in which the lengths of time series used for LE calculations should be chosen taking into
account the characteristic durations of periodic orbits embedded in chaos. The LEs presented
below were calculated using very long samples to ensure reliable averaging.
Further increase of Q restores the asymmetry of dynamical regimes due to the reverse cascade of
PD bifurcations which takes the system back to LC1:2. The AsLC1:2 is stable up to the high-Q
1:2TR seen in Fig. 20, where it then loses stability and the symmetric chaos is again the single
attractor of our system. An interesting qualitative phenomenon of the system is the continuous
strip of coexistence of the symmetric chaos and the LC1:2 formed by the end-of-chaos boundary
inside the 1:2TR for high-Q and inside the 1:2LP for low-Q. The coexistence of symmetrical
chaos with asymmetrical limit cycles is a remarkable peculiarity of the system (1). The
manifestation of this coexistence for the low-Q region is different from that described above for
the high-Q region. At the low-Q region bifurcation analysis discovered one more special chaotic
area bounded by the torus bifurcations of LC1:2 (see lines low-Q 1:2TR and 2:4TR in Fig. 20).
Consider the dynamics for the case α =2500. For Q-values beyond the LP1:2-line the system (1)
can be found in three coexisting states: LC1:2, LC3:6 (see below) and the symmetric chaos. The
evolution of the LC1:2 is presented in Fig. 22 where two LEs=0 starting at Q=0.542 indicates the
start of the torus, in agreement with the 1:2TR line in Fig. 20. The resulting torus is asymmetric,
and its development towards chaos is very limited for α=2500. Instead, the asymmetric branch
contains a sequence of asymmetric limit cycles: 5:10, 7:14, 9:18, 12:24, 13:26 and the system
demonstrates the transition to LC2:4 via 2:4TR bifurcation (Q=0.64) as prescribed by the map in
Fig. 20.
21
Fig.22. The dependence of LEs on Q (step=0.0005) for n=3, α =2500. The starting regimes are: LC1:2- solid lines
and the symmetric chaos – dashed lines. The insert is the LEs vs Q for regime LC3:6, which occurs at Q=0.518 and
overlaps with LC1:2 and chaos.
As for the symmetric chaos, for α=2500 it loses stability near the same Q-value where the LC1:2
lost stability, 1:2TR (Q=0.542), resulting in the system switching to an asymmetric weakly
chaotic regime originated from the period doubling cascade of the LC3:6 (insert in Fig.22). This
sudden loss of chaotic symmetry is illustrated in the time series in Fig. 23. Then the system
evolves up to the transition of the weak asymmetric chaos to asymmetric torus at Q=0.5457.
Fig. 23. The symmetric chaos loses stability and the system switches to asymmetrical weak chaos originated from
PD-cascade of LC3:6. α=2500, Q=0.5428
For larger values of α, the system demonstrates a wider area of stability for the symmetrical
chaos. For example, at α=3000 the symmetric chaos is stable up to the transition to LC2:4 (see
below the details of this transition) and it shares the phase space with LC1:2 and its asymmetric
derivatives. The evolutions of LEs for chaos and LC1:2 are presented in Fig. 24 where the new
regime LC3:6 is observed.
22
Fig. 24.Two sets of LEs vs Q for both regimes: start from LC1:2 – solid lines, start from chaos – dashed lines.
The LC3:6 is another exotic example of the coexistence of the symmetric chaos with an
asymmetric limit cycle (see Fig. 24). LC3:6 is the tripled version of LC1:2 (see Fig. 25a) and is
stable in a narrow band along the 3:6LP line (Fig. 20). It undergoes period-doubling bifurcation
to chaos, first to a weak asymmetric chaos followed by the transition to strong symmetric chaos
as directly observed in Fig. 25(b) and supported by the sharp increase of the first LE (black line
in Fig. 24 near Q=0.525). The LC3:6 is not a periodic window within the symmetric chaos.
Instead, it coexists with the symmetric chaos and with the LC1:2.
Fig. 25. (a)- the phase portrait of LC3:6, Q=0.507;(b)- the transition of weak asymmetric chaos to strong symmetric
one, Q=0.5249. α=3000.
Figure 24 shows that the chaos from the period-doubled LC3:6 matures and merges with the
symmetric chaos. The further evolution of this single chaotic attractor is shown in Fig. 26. The
sym-chaos loses stability, making a sharp transition to LC2:4. The period doubling of LC2:4
restores chaos, which is stable over the large interval of coupling strength up to the high-Q
period-halving cascade. The system sequentially crosses the regions with LC1:2, LC1:2+chaos,
23
pure chaos and torus, eventually reaching the TR-bifurcation back to stable anti-phase LC (see
Fig. 20).
Fig. 26. LEs vs Q starting from the symmetric chaos. α=3000.
4. Discussion
There are a huge number of papers concerning the collective behavior of nonlinear oscillators
published since Huygens's study of coupled clocks (see e.g. [1] and references there in). The
recent discovery of "chimeras" is an important and impressive example of the role of the
coupling type (see review [11]) in the formation of unexpected dynamical regimes. Although the
simple version of "quorum sensing" mechanism can be used to connect oscillators of different
nature [40, 41], it is most effective for the coupling of genetic oscillators imbedded in cognate
bacterial medium. The genetic networks are constructed from many nonlinear elements, each of
which may be engineered to accept the inter-oscillator signal produced in accordance with the
activity of other genes. The cooperation of this local property of genetic networks with the
indirect global coupling via signal molecule diffusion provides many possible combinations of
coupling schemes.
The design of our system, like that of the other versions of the model for coupled Repressilators
[27, 30-33], is the simplest one for 3-gene networks with one autoinducer. Being produced under
control of gene "b" promoter, the autoinducer activates the expression of downstream gene "c".
Despite linear diffusion of the autoinducer, the difference in gene "b" and "c" positions inside
the Repressilators' ring provides a completely possible, but a nonstandard coupling scheme. The
oscillations of repressors B1 and B2 concentrations are first converted to those of the signal
variables S1 and S2 with some time delay. Then, the signal molecules mix according to a quasi-
steady state approximation and transfer the impact of the superposition of delayed B1 and B2
oscillations onto the production rates of C1 and C2. As a result, the effective frequencies of
oscillations with the superposition, as well as their dependences on the model parameters, are
different from the internal frequencies of the Repressilators. As a result, the map in Fig. 27 over
24
the entire interval of α focuses on the major unexpected collective modes discovered and their
interrelations.
Fig. 27. Map of the major unexpected collective modes for the two identical QS-coupled ring oscillators. The paths
to chaos are presented in previous Sections.
The external boundary of the large parameter area with the rich set of attractors is formed by the
line of torus bifurcation of the anti-phase limit cycle. The appearance of torus in the system of
two coupled identical oscillators in the absence of an external periodic perturbation is a rare
event which requires specific choice of coupling term [42, 43]. The internal boundary of the
torus stability (ETR-line) up to α = 1000 is presented in the map Fig. 27 and it consists of two
parts separated approximately at α = 800. The loss of stability after the upper part of the ETR
line (800<α<1000) is a result of period doubling bifurcations of resonant cycles leading to the
chaotization of the system. Along the lower part of the boundary the torus is pushed from the
phase space by the LC5:5.
Although there are no indications of its bifurcation origin within the considered areas of
parameters, the LC5:5 is of principle importance for the dynamics below α=600. It is not a
resonant cycle limited by Neimark-Saker bifurcations. In contrast, up to α=600 it period-doubles
creating a very large area of chaos. For the larger α, its role in chaos production is diminished,
giving way to torus destruction as a path to chaos. Over the large interval above α=1000 a new
slightly asymmetrical LC9:9 is located within the closed area (Fig. 27, green line) and the chaos
is a result of its period doubling bifurcations. Omitting some details concerning the dynamics of
resonant cycles inside the low-Q region adjacent to the TR-line, we concentrate on the chaos
creation and conclude that in the system (1) the origin of symmetric chaos is the period doubling
bifurcations of different limit cycles: resonant cycles on torus, symmetric LC5:5 and asymmetric
25
LC9:9. Despite the identical nature of the coupled Represilators, the variability of return times
and amplitudes of variables in the 5:5 and As9:9 limit cycles are significant.
Further development of chaos over (α-Q)-plane demonstrates the surprising formation of
asymmetric limit cycles in narrow periodic windows, e.g. 3:4, 2:3, 3:5, which return to chaos via
period doubling bifurcations. The stability of chaos unexpectedly disappears (dashed black line
in Fig. 27) and the system jumps to the asymmetrical LC1:2 which is unlike standard limit cycles
within PWs. The LC1:2 (dark green line in Fig. 27) coexists with chaos and evolves
independently generating its own chaos via the period doubling bifurcations ending in the
restoration of the symmetrical chaos over a very large area of parameters. Although there are
known examples of chaos formation owing to the interaction of identical oscillators [44-47], the
indirect QS-coupling gives rise to new uncommon limit cycles being starting points for chaos
and its complex evolution over the parameter plane. Classical bifurcations of limit cycles --
period doubling and torus formation-- lead to standard processes of system chaotization.
However, under multistability conditions, they generate an unusual phase diagram with
domination of chaos and asymmetric attractors.
How general are the set of attractors and the observed structure of the map bearing in mind the
unavoidable deviations from perfectly identical oscillators and the reasonable shifts of other
model parameters, e.g. Hill coefficient n or the AI-induced activation rate κ? In previous
publications [32, 33] the unsuccessful choice of the main parameters revealed only the torus-
chaos transition corresponding to the lower part of the full map in Fig. 27. To check the
robustness of the detected attractors we calculated the boundaries of the basic regimes and found
that they changed on the map only in absolute coordinates but their relative positions remained
the same. In particular, the set n=3.15, k=10 shifts the boundaries of the regimes towards smaller
values of α but larger values of Q are then required to reveal all regimes (data not shown). A
small detuning of Repressilators via the parameter α mismatches did not remove the discovered
ways for chaos development and the appearance of hysteretic areas.
The formation of symmetric and asymmetric limit cycles inside quasiperiodic and chaotic
regimes, as well as the generation of extended regions with chaos, is a reliable consequence of
the specific choice of the QS equation used in system (1) for the coupling of the smooth ring
oscillators. The other ODE model of coupled Repressilators used in [27] is different from system
(1) by the location of the protein accompanying the production of autoinducer. However, the
other type of multistability manifested as the coexistence of in-phase and anti-phase regular and
complex oscillations is generated in extended areas of control parameters [28].
We speculate that the QS-dependent multistability may be realized in other synthetic models
with different versions of autoinducer production and its targets which may be important for the
understanding of the origin of variability, not only in genetic networks.
5. Acknowledgements
This work is partially supported by grants Russian Foundation for Basic Research 15-02-03236
and 17-01-00070. We thank Nikolay Sukhorukov for the calculations of Lyapunov exponents.
References
26
=======================================================
[1] Pikovsky, A, Rosenblum, M, Kurths, J. Synchronization A Universal Concept in Nonlinear Sciences,
Cambridge University Press; 2003.
[2] Pisarchik, AN, Feudel, U. Control of multistability, Physics Reports. 540 (2014) 167-218.
[3] Koseska, A, Volkov, E, Kurths, J. Oscillation quenching mechanisms: Amplitude vs. oscillation death,
Physics Reports. 531 (2013) 173-99.
[4] Taylor, AF, Tinsley, MR, Showalter, K. Insights into collective cell behaviour from populations of
coupled chemical oscillators, Physical Chemistry Chemical Physics. 17 (2015) 20047-55.
[5] Nagornov, R, Osipov, G, Komarov, M, Pikovsky, A, Shilnikov, A. Mixed-mode synchronization between
two inhibitory neurons with post-inhibitory rebound, Communications in Nonlinear Science and
Numerical Simulation. 36 (2016) 175-91.
[6] Prasad, A, Sakai, K, Hoshino, Y. Direct coupling: a possible strategy to control fruit production in
alternate bearing, Scientific Reports. 7 (2017) 39890.
[7] Kuramoto, Y, Battogtokh, D. Coexistence of Coherence and Incoherence in Nonlocally Coupled Phase
Oscillators, Nonlinear Phenomena in Complex Systems. 5 (2002) 380-5.
[8] Omel'chenko, OE, Wolfrum, M, Yanchuk, S, Maistrenko, YL, Sudakov, O. Stationary patterns of
coherence and incoherence in two-dimensional arrays of non-locally-coupled phase oscillators, Physical
Review E. 85 (2012) 036210.
[9] Tinsley, MR, Nkomo, S, Showalter, K. Chimera and phase-cluster states in populations of coupled
chemical oscillators, Nat Phys. 8 (2012) 662-5.
[10] Martens, EA, Thutupalli, S, Fourrière, A, Hallatschek, O. Chimera states in mechanical oscillator
networks, Proceedings of the National Academy of Sciences. 110 (2013) 10563-7.
[11] Panaggio, MJ, Abrams, DM. Chimera states: coexistence of coherence and incoherence in networks
of coupled oscillators, Nonlinearity. 28 (2015) R67.
[12] Zakharova, A, Kapeller, M, Schöll, E. Chimera Death: Symmetry Breaking in Dynamical Networks,
Physical Review Letters. 112 (2014) 154101.
[13] Purcell, O, Savery, NJ, Grierson, CS, di Bernardo, M. A comparative analysis of synthetic genetic
oscillators, Journal of The Royal Society Interface. 7 (2010) 1503-24.
[14] O'Brien, EL, Van Itallie, E, Bennett, MR. Modeling synthetic gene oscillators, Mathematical
Biosciences. 236 (2012) 1-15.
[15] Mandal, MK, Sarkar, BC. Ring oscillators: Characteristics and applications, Indian Journal of Pure and
Applied Physics. 48 (2010) 136-45.
[16] Elowitz, MB, Leibler, S. A synthetic oscillatory network of transcriptional regulators, Nature. 403
(2000) 335-8.
[17] Niederholtmeyer, H, Sun, ZZ, Hori, Y, Yeung, E, Verpoorte, A, Murray, RM, Maerkl, SJ. Rapid cell-free
forward engineering of novel genetic ring oscillators, eLife. 4 (2015) e09771.
[18] Potvin-Trottier, L, Lord, ND, Vinnicombe, G, Paulsson, J. Synchronous long-term oscillations in a
synthetic gene circuit, Nature. 538 (2016) 514-7.
[19] Gao, XJ, Elowitz, MB. Synthetic biology: Precision timing in a cell, Nature. 538 (2016) 462-3.
[20] Hellen, EH, Volkov, E, Kurths, J, Dana, SK. An Electronic Analog of Synthetic Genetic Networks, PLoS
ONE. 6 (2011) e23286.
[21] Hellen, EH, Kurths, J, Dana, SK. Electronic circuit analog of synthetic genetic networks: Revisited, Eur
Phys J Special Topics. 226 (2017) 1811-28.
[22] Garg, N, Manchanda, G, Kumar, A. Bacterial quorum sensing: circuits and applications, Antonie van
Leeuwenhoek. 105 (2014) 289-305.
27
[23] Smith, RP, Boudreau, L, You, L. Engineering Cell-to-Cell Communication to Explore Fundamental
Questions in Ecology and Evolution. in: SJ Hagen, (Ed.). The Physical Basis of Bacterial Quorum
Communication. Springer New York, New York, NY, 2015. pp. 227-47.
[24] Danino, T, Mondragon-Palomino, O, Tsimring, L, Hasty, J. A synchronized quorum of genetic clocks,
Nature. 463 (2010) 326-30.
[25] McMillen, D, Kopell, N, Hasty, J, Collins, JJ. Synchronizing genetic relaxation oscillators by intercell
signaling, Proceedings of the National Academy of Sciences. 99 (2002) 679-84.
[26] Kuznetsov, A, Kaern, M, Kopell, N. Synchrony in a Population of Hysteresis-Based Genetic Oscillators,
SIAM Journal on Applied Mathematics. 65 (2004) 392-425.
[27] Garcia-Ojalvo, J, Elowitz, MB, Strogatz, SH. Modeling a synthetic multicellular clock: Repressilators
coupled by quorum sensing, Proceedings of the National Academy of Sciences of the United States of
America. 101 (2004) 10955-60.
[28] Potapov, I, Volkov, E, Kuznetsov, A. Dynamics of coupled repressilators: The role of mRNA kinetics
and transcription cooperativity, Physical Review E. 83 (2011) 031901.
[29] Li, C, Chen, L, Aihara, K. Stochastic synchronization of genetic oscillator networks, BMC Systems
Biology. 1 (2007) 6.
[30] Yi, QZ, Zhang, JJ, Yuan, ZJ, Zhou, TS. Collective dynamics of genetic oscillators with cell-to-cell
communication: a study case of signal integration, The European Physical Journal B. 75 (2010) 365-72.
[31] Yi, Q, Zhou, T. Communication-induced multistability and multirhythmicity in a synthetic
multicellular system, Physical Review E. 83 (2011) 051907.
[32] Ullner, E, Koseska, A, Kurths, J, Volkov, E, Kantz, H, García-Ojalvo, J. Multistability of synthetic
genetic networks with repressive cell-to-cell communication, Physical Review E. 78 (2008) 031904.
[33] Ullner, E, Zaikin, A, Volkov, EI, García-Ojalvo, J. Multistability and Clustering in a Population of
Synthetic Genetic Oscillators via Phase-Repulsive Cell-to-Cell Communication, Physical Review Letters.
99 (2007) 148103.
[34] Hellen, EH, Volkov, E. Flexible dynamics of two quorum-sensing coupled repressilators, Physical
Review E. 95 (2017) 022408.
[35] Brewster, RC, Jones, DL, Phillips, R. Tuning Promoter Strength through RNA Polymerase Binding Site
Design in Escherichia coli, PLOS Computational Biology. 8 (2012) e1002811.
[36] Davis, JH, Rubin, AJ, Sauer, RT. Design, construction and characterization of a set of insulated
bacterial promoters, Nucleic Acids Research. 39 (2011) 1131-41.
[37] Hellen, EH, Dana, SK, Zhurov, B, Volkov, E. Electronic Implementation of a Repressilator with
Quorum Sensing Feedback, PLoS ONE. 8 (2013) e62997.
[38] Ermentrout, B. Simulating, Analyzing, and Animating Dynamical Systems: A Guide to XPPAUT for
Researchers and Students. Philadelphia, PA, SIAM; 2002.
[39] Doedel, EJ, Fairgrieve, TF, Sandstede, B, Champneys, AR, Kuznetsov, YA, Wang, X. AUTO-07P:
Continuation and bifurcation software for ordinary differential equations. 2007.
[40] Taylor, AF, Tinsley, MR, Wang, F, Huang, Z, Showalter, K. Dynamical Quorum Sensing and
Synchronization in Large Populations of Chemical Oscillators, Science. 323 (2009) 614-7.
[41] De Monte, S, d'Ovidio, F, Danø, S, Sørensen, PG. Dynamical quorum sensing: Population density
encoded in cellular dynamics, Proceedings of the National Academy of Sciences. 104 (2007) 18377-81.
[42] Bondarenko, VE, Cymbalyuk, GS, Patel, G, DeWeerth, SP, Calabrese, RL. Bifurcation of synchronous
oscillations into torus in a system of two reciprocally inhibitory silicon neurons: Experimental
observation and modeling, Chaos: An Interdisciplinary Journal of Nonlinear Science. 14 (2004) 995-1003.
[43] Rosenblum, M, Pikovsky, A. Self-Organized Quasiperiodicity in Oscillator Ensembles with Global
Nonlinear Coupling, Physical Review Letters. 98 (2007) 064101.
28
[44] Schreiber, I, Marek, M. Transition to chaos via two-torus in coupled reaction-diffusion cells, Physics
Letters A. 91 (1982) 263-6.
[45] Sporns, O, Roth, S, Seelig, FF. Chaotic dynamics of two coupled biochemical oscillators, Physica D:
Nonlinear Phenomena. 26 (1987) 215-24.
[46] Hakim, V, Rappel, W-J. Dynamics of the globally coupled complex Ginzburg-Landau equation,
Physical Review A. 46 (1992) R7347-R50.
[47] Nakagawa, N, Kuramoto, Y. Collective Chaos in a Population of Globally Coupled Oscillators,
Progress of Theoretical Physics. 89 (1993) 313-23.
|
1302.1641 | 1 | 1302 | 2013-02-07T04:40:52 | Membrane mediated aggregation of curvature inducing nematogens and membrane tubulation | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.SC"
] | The shapes of cell membranes are largely regulated by membrane associated, curvature active, proteins. We use a numerical model of the membrane with elongated membrane inclusions, recently developed by us, which posses spontaneous directional curvatures that could be different along and perpendicular to its long axis. We show that, due to membrane mediated interactions these curvature inducing membrane nematogens can oligomerize spontaneously, even at low concentrations, and change the local shape of the membrane. We demonstrate that for a large group of such inclusions, where the two spontaneous curvatures have equal sign, the tubular conformation and sometime the sheet conformation of the membrane are the common equilibrium shapes. We elucidate the factors necessary for the formation of these {\it protein lattices}. Furthermore, the elastic properties of the tubes, like their compressional stiffness and persistence length are calculated. Finally, we discuss the possible role of nematic disclination in capping and branching of the tubular membranes. | physics.bio-ph | physics | Membrane mediated aggregation of curvature inducing nematogens and
http://dx.doi.org/10.1016/j.bpj.2012.12.045, Biophysical Journal
membrane tubulation
N. Ramakrishnan∗ and P. B. Sunil Kumar†
Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India
John H. Ipsen‡
MEMPHYS- Center for Biomembrane Physics, Department of Physics and Chemistry,
University of Southern Denmark, Campusvej 55, DK-5230 Odense M, Denmark
(Dated: September 30, 2018)
Abstract
The shapes of cell membranes are largely regulated by membrane associated, curvature active, proteins. We use
a numerical model of the membrane with elongated membrane inclusions, recently developed by us, which posses
spontaneous directional curvatures that could be different along and perpendicular to its long axis. We show
that, due to membrane mediated interactions these curvature inducing membrane nematogens can oligomerize
spontaneously, even at low concentrations, and change the local shape of the membrane. We demonstrate that for
a large group of such inclusions, where the two spontaneous curvatures have equal sign, the tubular conformation
and sometime the sheet conformation of the membrane are the common equilibrium shapes. We elucidate the
factors necessary for the formation of these protein lattices. Furthermore, the elastic properties of the tubes, like
their compressional stiffness and persistence length are calculated. Finally, we discuss the possible role of nematic
disclination in capping and branching of the tubular membranes.
PACS numbers: PACS-87.16.D-, Membranes, bilayers and vesicles. PACS-05.40.-a Fluctuation phenomena, random pro-
cesses, noise and Brownian motion. PACS-05.70.Np Interfaces and surface thermodynamics
3
1
0
2
b
e
F
7
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
1
4
6
1
.
2
0
3
1
:
v
i
X
r
a
∗Electronic address: [email protected]
†Electronic address: [email protected]
‡Electronic address: [email protected]
1
I.
INTRODUCTION
Membrane shape deformations are key phenomena in a multitude of cellular processes, including pro-
tein sorting, protein transport, organelle biogenesis and signaling.
In the last decade a profusion of
regulatory proteins facilitating such shape changes of the cellular membranes has been unravelled, with
the BAR protein superfamily [1], the Pex11 family [2] and coat proteins [3] as notable examples. The
possibility of such mechanisms has long been anticipated in the biophysical literature[4, 5]. However
the experimental and theoretical difficulties involved have hampered the establishment of a quantitative
basis for interpreting such phenomena in cell biology. Recently, we had overcome one such obstacle by
the establishment of a computer simulation technique to study how the cooperative effects of membrane
inclusions, imposing a curvature along the direction of its orientation, remodels vesicular membranes[6].
In this work we aim at describing, from a theoretical point of view, the effect of a large group of these
membrane curving proteins, which can be considered as effectively elongated objects in the plane of the
membrane. We consider inclusions with approximate π-symmetry, i.e. the protein can be considered
as essentially indistinguishable from its form rotated by 180◦ around the protein center in the plane
of the membrane. The membrane inclusions we consider, has thus some similarity with nematogens
in 3-D nematic liquid crystals. However, they are embedded in a membrane and may couple to its
geometry, and it is only the part of the protein in contact with the membrane, which will be subject
to these symmetry requirement. Therefore, we cannot consider these membrane inclusions as simple
liquid crystal nematogens restricted to Euclidean two dimensional surfaces. In the following we will refer
to such membrane inclusions as membrane nematogens. Large groups of membrane curving proteins
fall into this category of membrane nematogens. An example is the BAR proteins(proteins containing
both BAR domains and/or N-terminal helices), where both the N-terminal amphipatic helices and the
banana-shaped, positively charged, dimeric interface with the membrane, induces directional curvature
[7 -- 12]. The caveolin protein family[13], which form dimers and are bound to the membrane by a pair
of hairpins and the reticulon, DP1 and Yop1p involved in the formation of smooth ER [15, 16], and
are anchored to the membrane by two similar hairpins are also examples. The cell biology literature
has provided good evidence for that the insertion of amphipathic helical peptide sequences not only
provide a binding mechanism, but also gives rise to local modulation of the membrane curvature [17, 18].
More solid, quantitative support for this conjecture is given from biophysical experiments [19] and theory
[20] . Furthermore, biophysical studies has demonstrated that curvature active membrane inclusions
have dramatic effects on the cooperative behavior with a complex interplay between lateral ordering and
membrane shape. However, the detailed mechanisms leading to the specific complex membrane-protein
structures have not been characterized. This work will elucidate some aspects of these mechanisms for
the membrane nematogens.
Some of the key processes involved in the structural organization of membrane nematogens described in
the cell biology literature, can be categorized as follows: (1) the aggregation of the nematogens - the pro-
cess where membrane proteins upon activation and/or binding to the membrane spontaneously aggregate
and form functional cluster of proteins in the membrane[9 -- 11, 14], (2) Tubulation of membranes, where
the aggregate and the membrane develop tube-like membrane structures (e.g. sorting endosomes[21, 22]
and Mitochondrial outer membrane[23], formation of T-tubules in Drosophila[24]) and (3) The formation
of protein lattices, often characterized by helical arrangement of the proteins spiraling around the tubular
membrane, e.g. for dynamin [25, 26] or caveolin [27].
In this work, we will demonstrate by numerical analysis of a possible physical model, which captures
the membrane conformations and the organization of in-plane nematogens, that the above mentioned
processes directly results from the cooperative thermodynamic behavior of the nematogens coupled to
the flexible membrane. Also, we will discuss aspects of the stability of membrane tubes and the formation
of the edges for membrane sheets. Our model gives a coarse description of the membrane, which capture
properties of the membrane which are essential for its large scale organization. Despite the simplicity of
the model, the parameter space is too large for a comprehensive discussion of it's phase behavior. Rather,
we will focus on some generic features of the model which may well give a framework for interpreting the
experimental observations of cellular membrane morphogenesis. Previously, protein induced membrane
tube formation has been considered by a phenomenological model involving scalar fields [28], and the
coupling between membranes and inclusions with directional curvature was modelled in [29 -- 32].
The paper is organized as follows: In Section II the physical model of the interacting system of mem-
brane and membrane nematogens are presented, while details about the numerical analysis is given in
Supplementary Materials. Section III on Results and Discusion present some generic properties of
2
the model and discuss their possible relevance to experimental results. In section III A-III C the aggrega-
tion of proteins and membrane domain formation, membrane tubulation and formation of protein lattices
are described in the framework of the model. Section III D discusses the elastic properties of membrane
tubes and their relevance to observable effects. Much of the characterization of the elasticity of protein
lattices is based on a continuum version of the model discussed Supplementary Materials. Section
III E discusses mechanisms of closing, capping and branching of membrane tubules and the possible role
of nematic point defects. Section III F describes aspects of the stability of membrane tubules with more
membrane curvature components. In section III G the interplay between sheet and tubule formation is
described and possible implications for cell organel morphology is discussed. Some perspective on the
modeling of membrane morphogenesis is given in Conclusion, Section IV. We will in this work specialize
to properties of membranes with inclusions which posses directional curvatures of equal sign. We will
consider cases with different signs of the directional curvatures in a seperate publication.
II. MODEL
The modeling of the effects of in-plane nematogens on membrane structure, will in this work, be
treated with a discretized description of the surface as a randomly triangulated mesh. A continuous
surface conformation is approximated by a collection of triangles glued together to form a closed surface
of well defined topology. A triangulated surface, with spherical topology, thus consist of N vertices
connected by NL = 3(N − 2) links, which enclose NT = 2(N − 2) triangles. Each vertex v is assigned
a position (cid:126)Xv. This tesselatations of the surface form the basis for a coarse grained description of the
membrane, where only the gross features of the structure and interactions are important.
FIG. 1:
(a) A one ring triangulated patch around a vertex v. The shaded region represents the tangent plane
at v and N (v) its corresponding normal. c1v and c2v are the maximum and minimum principal curvatures,
respectively, along principal directions t1v and t2v. (b) Illustration of the nematic field vector n defined on the
tangent plane of vertex v. (c) A vesicle of spherical topology with spatially random surface nematogens.
The triangulation and the vertex position form together a discretized surface, a patch of which is given
in fig:??. The geometry of the continuous surface, which is approximated by the triangulated surface,
can now be characterized by a number of surface quantifiers, e.g. the curvature tensor, the principal
directions (t1v & t2v), the corresponding principal curvatures (c1v & c2v) and surface normal, N (v), at
each vertex v. The details can be found in [6]. The discretized Helfrich's free energy[34] can then be
evaluated as
N(cid:88)
v=1
Hc =
κ
2
Av
3
H 2
v
(1)
where Hv = (c1v + c2v)/2 the mean curvature at vertex v and Av is the area of the surface patch occupied
by the triangles adjacent to vertex v. κ is the bending rigidity of the membrane. Furthermore we are in
a position to calculate the directional curvatures along and perpendicular to a unit vector n along the
surface by use of Gauss formula:
Hv,(cid:107) = c1v cos2 ϕv + c2v sin2 ϕv,
Hv,⊥ = c1v sin2 ϕv + c2v cos2 ϕv,
3
(2)
planetangentvN(v)c1vc2vt2vt1vvn(v)ϕvt2vt1v(a)(b)(c) Mixed phasedemixing and constrictionActivation of field Bwhere ϕv is the angle between n and the principal direction t1v. Such an orientational spontaneous
curvature may be induced by a membrane nematogen with an orientation in the plane of the membrane
given by n. In addition to the interaction with the membrane, nematogens may tend to orient along
each other at close proximity due to steric, electrostatic and dispersion interactions [35]. In the present
study, we focus only on the two dimensional orientational interactions promoted by the underlying, non-
planar, fluctuating membrane [36 -- 40]. The π-symmetry of the individual nematogens dictates that the
simplest form of their self interaction should be of the type cos2(θvu) and sin2(θvu), where θvu is the angle
between nv and nu at neighboring vertices. We choose to represent the interactions between membrane
nematogens by an extension of the well-established Lebwohl-Lasher model of nematic ordering in presence
of vacancies, here implemented on a triangulated surface model of a membrane.
The nearest neighbor interaction between the nematogens is composed of an isotropic component
represented by an interaction strength J and an anistropic (quadrupolar) correction measured by the
interaction constant LL. The total interaction between the membrane nematogens thus takes the form
(cid:26)
(cid:88)
(cid:104)vu(cid:105)
(cid:18) 3
2
(cid:19)(cid:27)
Hfield =
J
2 − LL
−
cos2(θvu) −
1
2
IvIu,
(3)
where the sum is over nearest neighbour vertices. Iv = 0, 1 is an occupation variable, which is unity if
vertex "v" is occupied by a nematogen and zero if otherwise. The calculation of the θvu is non-trivial,
since the angle between spatially separated nematogens are measured after the parallel transport of vec-
tors along the curved surface [6]. With this measure of the angular differences, Eq.(3) models the in-plane
interactions of the nematogens mediated by the membrane. The direct distance dependent interactions
through the cytosol is not taken into account in this model of membrane-protein conformations. Suf-
ficiently large, positive LL favors in-plane ordering of the nematogens. The effect on the anisotropic
elasticity of the membrane due to the nematogens, in analogy with the discretized Helfrich free energy,
takes the form [6]:
(cid:16)
(cid:26) κ(cid:107)
N(cid:88)
2
v=1
Hnc =
(cid:17)2
Hv,(cid:107) − H 0(cid:107)
+
κ⊥
2
(cid:27)
(cid:0)Hv,⊥ − H 0⊥(cid:1)2
IvAv
(4)
H 0(cid:107) and H 0⊥ are the spontaneous curvatures along n and n⊥, while κ(cid:107) and κ⊥ are the corresponding
directional membrane bending elastic constants.
Self avoidance of the discretized surface is ensured by imposing constraints on the neighboring vertex
distance and on the dihedral angles between neighboring faces[6]. The equilibrium properties of the
discretized surface can now be evaluated by standard Monte Carlo sampling of Boltzmann's probability
distribution ∼ exp
at fixed concentrations of the membrane nematogens.
A general description of the implemetation of such numerical models and further details about the
simulations are given in [6].
− 1
kB T [Hc + Hfield + Hnc]
(cid:16)
(cid:17)
Finally, we will make some considerations about length scales. The lattice model is a highly coarse-
grained representation of the membrane, designed to capture the large length-scale properties of mem-
branes with inclusions. Therefore, the triangulated surface represent a collection of membrane patches
with a characteristic length scale. A natural choice of length scale is to identify a tether length with the
lateral extension of a membrane inclusion. Some examples here are CIP4 F-BAR with a length of 22
nm[41] or dynamin which extent about 25 nm [25].
The computer simulations of the discrete model provide us with insight into the nature of equilibrium
configurations for a choice of model parameters. To complement the numerical simulations, it is useful
to consider the corresponding continuum model in the limit of membrane nematogen with 100% surface
coverage. It is an extension of Helfrich's bending free energy functional [42, 43]
(cid:73)
(cid:26) KA
F =
dA
Tr (∇n : ∇n) +
2
+
κ(cid:107)
2
(Hn,(cid:107) − H 0(cid:107) )2 +
κ
2
(2H)2
(Hn,⊥ − H 0⊥)2(cid:111)
κ⊥
2
√
where KA = 3
3
2 LL. In Supplementary Material is presented such an analysis of the mechanical
properties of a tubular membrane with a protein coat and analytical expressions reflecting, tube radius,
persistence length and protein organization are also derived.
4
III. RESULTS AND DISCUSSION
In this section we will present some key aspects resulting from the coupling of membrane nematogen
proteins to lipid membranes. It will both contain results from computer simulations of the aforementioned
model, which are non-perturbative, along with theoretical analysis of the continuum model, of a more
perturbative character to qualify the numerical finding. Throughout the discussion the parameter LL
has a relatively high value (several kBT in a range where nematic ordering is favored). Furthermore, the
implications of our results on the experimental systems in vivo and in vitro will be discussed.
A. Aggregation and membrane domain formation of membrane nematogens
A common feature of membrane nematogens is their strong tendency to self-associate, driven by the
flexible geometry of the membrane - in this manuscript, we call this self associated structure to be an
aggregate or a domain. Self association has been observed for a wide range of model parameters κ(cid:107),
κ⊥, H 0(cid:107) and H 0⊥. All results presented in the following corresponds to system size with N = 2030
vertices. When the fraction of nematogens φA = 0.3, LL = 3 and J = 0 (in units of kBT ), the flexible
membrane with curvature coupled to the nematic orientation, gives rise to co-existence of nematically
ordered domains and the isotropic dilute phase, this is shown in figure-2(b). This is to be compared with
the planar Lebwohl-Lasher model on a random triangular lattice, at the same concentration, where the
isotropic phase is stable (see Supplementary Materials). Additional direct repulsive interactions J ≤ −0.5
between the membrane nematogens can reestablish the isotropc phase, which is shown in Fig. 2(a).
The aggregation of membrane nematogens cause shape deformation of the whole membrane with the
collective involvement of all the degrees of freedoms, lateral orientation, lateral position and membrane
conformation.
In general the lateral domain formation depends on all the involved parameters - e.g.
increasing J promotes the aggregation and can change the aggregate shape as shown in Fig. (2)(c).
FIG. 2: Equilibrium membrane conformations, with φA = 0.3, κ = 20, κ(cid:107) = 5 and H 0(cid:107) = 0.5 and LL = 3, for
different range of J.
The effect of concentration is shown in Fig.(3) for surface coverage in the range φA = 0.1 − 0.7, which
display a series of complex shape deformations connected to different aggregate structures. More details
will be discussed in section III G.
The aggregation of membrane nematogens also has a temporal aspect. In Fig.(4) we have shown a
Monte Carlo time series, for a membrane coverage of 10% nematogens, to illustrate the sequence of
domain formation and membrane curvature induced changes leading to the equilibrium structure. The
membrane nematogens in an initial randomly dispersed orientation assemble into smaller orientationally
ordered domains mediating the final equilibrium structure. These ordered domains often appear as
metastable configurations, which either disperse again due to lateral fluctuations or they will eventually
funnel into a equilibrium domain configuration. The Monte Carlo dynamics does not reflect the physical
kinetics very well, but is useful in identifying kinetic paths connecting various metastable states that lead
to the global minima [44, 45] .
The aggregation of membrane inclusions mediated by membrane curvature deformations and fluctuations
is not specific for nematogens, but is a more general phenomena for membrane curvature active com-
ponents. It has been well understood in the framework of models for curvature instabilities [5, 19, 46],
5
(c) Tubular(a) Spherical (b) Disc−1<J<−0.5−0.5<J<0.50.5<J<1FIG. 3: Equilibrium configurations for varying composition. κ = 10, κ(cid:107) = 5, κ⊥ = 0, H 0(cid:107) = 0.5, H 0⊥ = 0, J = 0,
φA = 0.1 − 0.7 and LL = 3.
FIG. 4: Aggregation of membrane inclusions for κ = 10, κ(cid:107) = 5, κ⊥ = 0, H 0(cid:107) = 0.5, H 0⊥ = 0, J = 0, φA = 0.1
and LL = 3. Monte Carlo time series showing a) random initial configuration of membrane nematogens, b)
intermediate state with multiple nematic domains and c) equilibrium conformation where all the small domains
coarsen into a single patch.
and has also been demonstrated that simple amphiphathic inclusions, e.g. antimicrobial peptides like
Magainin or Melittin [19, 47] and viral membrane active proteins like NSB4 of Hepetites C[48].
The self-association of these membrane components thus needs not to be facilitated by strong direct
attractive interaction amongst them. The coupling to the membrane geometry provides additional indirect
membrane conformation mediated attractive forces making them to slip into bound structures involving
both the proteins and membrane curvature. However, the structure of the aggregates are dependent on
the details of the molecular structure and the direct interactions. A general feature of these aggregates is
that they appear as nematically ordered domains, where the nematogens form elongated oriented patches
with well defined curvature characteristics, e.g. ridges or cylindrical rims. In the following we will in
particular focus on the tube-like structures.
6
(a)φA=0.1(b)φA=0.2(c)φA=0.3(d)φA=0.4(e)φA=0.5(f)φA=0.6(g)φA=0.7(a)TMCS=0(b)TMCS=1.6×105(c)TMCS=17.5×105B. Tube formation
The most prevalent equilibrium domain structure is the nematic tube, where the membrane protrude
into a cylinder like configuration with the membrane nematogens forming a coat around the cylinder.
Also for tube formation the overall interaction strength (J) between the membrane nematogens plays a
secondary role. It's most pronounced effect is to widen the concentration range for tubulation and to
enhance the line tension at the domain boundary, which can induce fission of tubes by narrowing the tube
at the boundary of the domain, as in Fig(2(c)). The effect of concentration of membrane nematogens on
the membrane tubulation phenomena is shown in Fig.(3(g)). For large concentrations of nematogens or
increasing values of J the tubes are the characteristic equilibrium structures shown in Figs. (2,3).
The radius of the equilibrium membrane tubes appears to be relatively well-defined. The radius of the
tube with nematic order can be calculated on basis of the continuum model Eq.(5) for the chosen model
parameters (see Supplementary Material):
1H0(cid:107)
1H0⊥
¯r =
(cid:113) κ(cid:107)+κ
(cid:113) κ⊥+κ
κ(cid:107)
κ⊥
for κ⊥ = 0
for κ(cid:107) = 0
(5)
So, the radius ¯r is set by the curvature elastic model parameters, involving the absolute value of the
directional spontanous curvatures, modulated by the curvature elastic constants. It follows from Eq.(5)
that the actual tube radius is somewhat larger than the inverse directional spontaneous curvatures and
dependent on the relative strength of the elastic constants.
In experimental systems the membrane tube dimensions can vary considerably with different types
of proteins in the cell[1]. Membrane tubes formed in vitro by curvature active proteins also display a
considerable variability in size. Frost et al. [41] have studied the effect of a number of mutants of CIP4
F-BAR on liposomes. By mutations they find a big variations of tube diameters in the range of 50 to
100 nm.
Membrane tubes induced by membrane inclusions are common phenomena in biological cells, both
as more static structures like T-tubules of the muscle cells[24] or more temporal structures like sorting
endosomes [49]. The examples shown in Fig.(2,3) corresponds to the cases where spontaneous curvatures
are positive, like that induced when F- BAR-domain proteins bind to organelle membranes. However, if
the proteins induce negative spontanous curvatures, as in I-BAR domain proteins, it gives rise to tubular
invaginations as shown in Fig.(5) .
As can be seen from Fig.(5), for proteins with large negative spontaneous directional curvatures, at low
concentration (φA = 0.1 − 0.3), we obtain tubes growing into the interior of the vesicle. As φA increases,
tubes disappear and saddle like regions appear. The inner tubes and saddle like regions coexist again for
large concentrations φA > 0.8.
C. Protein lattices
Membrane nematogens organize as nematically ordered domains and coat around the membrane to
form tubes. Nematogens orient perpendicular to the tube axis when κ⊥ = 0, κ(cid:107) (cid:54)= 0 and H 0(cid:107) > 0.
Similarly, κ(cid:107) = 0, κ⊥ (cid:54)= 0 and H 0⊥ (cid:54)= 0 leads to an arrangement of the nematogens along the tube
direction. For the common membrane nematogen both these parameters are non-vanishing. Such a case
is shown in Fig.(6).
The helical arrangement of the membrane nematogens at the tube surface can be easily understood
considering that in general such arrangement will give rise to a global nematic ordering of the membrane
nematogens (generalized spirals are the only geodesic curves on long tubes) and the radius is set by the
elastic terms. The coupled expressions for the mean values for tube radius ¯r and the angle ¯ϕ between the
tube direction and the nematogen orientation, for different regimes of the dimension less parameter ψ:
7
FIG. 5: Equilibrium configurations for vesicle with negative spontanous curvatures. κ = 10, κ(cid:107) = 5, κ⊥ = 0,
H 0(cid:107) = −0.5, H 0⊥ = 0.0, J = 0 and φA = 0.1 − 0.7.
(cid:115)
(cid:118)(cid:117)(cid:117)(cid:116)(cid:0) κ
(cid:115)
¯r =
κ⊥ + κ
κ(cid:107)(H 0(cid:107) )2 + κ⊥(H 0⊥)2
(cid:1)(cid:0)κ(cid:107) + κ⊥(cid:1) + κ⊥κ(cid:107)
(cid:16)
(cid:17)2
2
κ⊥κ(cid:107)
H 0(cid:107) + H 0⊥
κ(cid:107) + κ
κ(cid:107)(H 0(cid:107) )2 + κ⊥(H 0⊥)2
for
ψ ≤ 0,
0 < ψ < 1
ψ ≥ 1
The parameter ψ is given by the model parameters as:
(cid:115)
1 +
(cid:18) 1
κ(cid:107)
κ
2
(cid:19)
+
κ⊥
κ(cid:107)
+
1
κ⊥
ψ =
κ(cid:107)H 0(cid:107) − κ⊥H 0⊥
H 0(cid:107) + H 0⊥(κ(cid:107) + κ⊥)
similarly for the angle ¯ϕ:
0
(cid:16)
1
(cid:17)
κ(cid:107)H 0(cid:107) − κ⊥H 0⊥
κ⊥ + κ(cid:107)
for
ψ ≤ 0,
¯r + κ⊥
0 < ψ < 1
ψ ≥ 1
cos2( ¯ϕ) =
(6)
(7)
(8)
So, in general both ¯ϕ and ¯r are set by the model parameters. A derivation of the above expressions are
given in Supplementary Materials .
8
(a)φA=0.1(b)φA=0.2(c)φA=0.3(d)φA=0.4(e)φA=0.5(f)φA=0.6(g)φA=0.7FIG. 6: Protein lattices. Modes of a tubular membrane at different state points, for κ = 10 and LL = 3. (a)
Tubular conformation with (cid:104)ϕ(cid:105) = 0 for κ(cid:107) = 5, κ⊥ = 0, H 0(cid:107) = 0.4, (b) Spiral modes of the tube with (cid:104)ϕ(cid:105) = 0
seen for κ(cid:107) = 5, κ⊥ = 0, H 0(cid:107) = 0.4, and (c) rearrangement of nematics into spiral modes ((cid:104)ϕ(cid:105) (cid:54)= 0) when κ(cid:107) = 5,
κ⊥ = 5, H 0(cid:107) = 0.4, H 0⊥ = 0.25
The spiral organization of the membrane coating proteins has now been observed for many tubular
membrabe systems in vivo and in vitro, e.g for the F-BAR proteins[26, 41]. EM-tomographs of tubules of
CIP4 F-BAR on liposomes [41] show a fairly dense packing arrangement in the helical tube. The average
tube diameter is around 68 nm and the helical angle is about ϕ = 40◦. Such arrangements are termed
protein lattices in the cell biology literature. It is found that the helical angle ϕ of the protein lattice with
respect to the tube direction adjusts to the tube diameter such that the directional curvature is about the
same. For a similar type of experiment with dynamin[50] the membrane tubes of radius r (cid:39) 23nm with
densely packed helical dynamin coat was observed with a helical pitch of about 15nm (corresponding to
a helical angle ϕ = 80◦).
Our simulation results suggests that the spiral organization of the protein coat on the tube need not
be a result of polymerization as often referred in the literature, but can be a self-assembly process of
the curvature active proteins mediated by the membrane. Furthermore, the modelling suggest that these
protein lattices are not conventional two dimensional lattice structures like polymerized membranes or
graphene, but rather two dimensional nematic liquid crystalline structures. In the model there is no terms
which can distinguish between a right or left turning helix, i.e. the helical arrangement is the result of
a spontaneous symmetry breaking. However, the smallest chiral symmetry breaking contribution to the
free energy can favor one of the helical orientations without having an effect on any other parameters.
D. Thermal stability of membrane tubes
While our model parameter determine the mean physical properties of the tubes , e.g. the radius, we
expect the tubular membranes to display an elastic response to deformations in its shape and organization
of the membrane nematogens. This can, for e.g., be reflected in the variation of the shape characteristics
due to thermal fluctuations. For analysis of such deformations the continuum description of coated
9
(a)(b)(c)membrane tubes are suitable and the details can be found in Supplementary Material. It is shown
that in general the deviations in the orientation of the membrane nematogen and the tube radius are
strongly correlated. The thermally induced fluctuations in the radius is found to be
(cid:104)(δr)2(cid:105)
¯r2 =
kBT
4π
κ(cid:107) + κ⊥
κ(cid:107)κ⊥ + (κ(cid:107) + κ⊥) κ
2
for 0 < ¯ψ < 1,
(9)
where ¯r and ¯ϕ are respectively the equilibrium tube radius and nematic orientations and ¯ψ = cos2 ¯ϕ.
We note that the relative variance in r has an upper limit kB T
2πκ . With a typical range κ ∼ 20 − 50kBT
this ratio in Eq.(9) is of the order 0.01. For CIP4 F-BAR, reconstituted on liposomes, cryo-tomography
measurements give ¯r = 33nm and (cid:104)(δr)2(cid:105)/¯r2 (cid:39) 0.01 [41]. If this observed variation in tube thickness
is interpreted as frozen in thermal variations, it is in agreement with the the above theory. For rigid
membranes with large κ and/or large κ(cid:107), κ⊥ values we can consider the thermaly excited variations in r
as small. Similarly, we can estimate the thermal fluctuations around cos2( ¯ϕ) for such a segment as,
(cid:104)(δ cos2(ϕ))2(cid:105) =
kBT
4π
κ(cid:107)ψ2 + κ(cid:107)(1 − ψ)2 + κ
κ(cid:107)κ⊥ + (κ(cid:107) + κ⊥) κ
2
2
for 0 < ¯ψ < 1.
(10)
To our knowledge no experimental reports on the random variations in the helical angle has been given.
A third type of deformation to consider is the bending of the tubes. It is shown in Supplementary
Materials that when r is a constant along the tube, the free energy expression is relatively simple. In
particular we find that the free energy of bending for a tubular membrane takes the approximate form,
dsλ(s)2,
(11)
(cid:90) L
0
∆Ftot ≈
1
2
kBT lP
π¯r(cid:0)KA + κ + κ(cid:107)(1 − ¯ψ)2 + κ⊥ ¯ψ2(cid:1)
kBT
where s is the the arc length and λ(s) is the line curvature along the tube, while lP is the persistence
length of the tube:
lP =
.
(12)
There are few experimental measurements of the persistence length of membrane tubes with protein
lattices. For the F-BAR FBP17 producing tubes of radius r(F BP 17) = 34nm the persistence length was
measured to lP (F BP 17) = 142µm [41] while for amphiphysin r(amph) ∼ 7nm[51] and lP (amph) = 9µm
while for dynamin r(dynamin) (cid:39) 20nm and lP (dynamin) = 37µm. A calculations of lP from Eq.(12)
solely based on κ gives predictions which are an order of magnitude too small, which indicates that other
elastic constants κ(cid:107), κ⊥ and KA gives the main contributions to lP .
E. Capping the tubes, Defects
The formation of membrane tubes with helical coats seems to be generic for systems with membranes
with membrane nematogens. Either the helical coat has to terminate resulting in an interfacial curve
separating the coated and uncoated regions or the vesicle should sprout tubes and buds with the tips
having a pair of point defects. The way this takes place in the tube end or at a domain boundary is
mainly determined by the competition between interfacial tension, which in our model is largely regulated
by the parameter J, and bending modulus. In Fig(7) is shown that when the interaction parameter J
is increased the interfacial line shrinks, first transforming the vesicle from a disk to a structure with
partially coated tubes and buds, but still no defects. Further increase in the line tension will result in
tubes and buds that are fully coated but minimizing the length of the interfacial line between coated and
uncoated regions. It does so by either moving the interfaces to the end of the tube forming a pair of
point defects or deforming the membrane to form a narrow neck. Note that the line does not shrink to a
single point defect of strength +1 but instead forms a pair of +1/2 defects bound to each other. This is
the result of of the π symmetry of the membrane nematogen and the strong coupling between membrane
curvature and nematic orientation [52 -- 54]. To our knowledge no details about the capping of the coated
membrane tubes have been provided by experiments.
10
FIG. 7: A partly decorated membrane with κ = 20, κ(cid:107) = 5, LL = 3. Shown are: (a) a disc without a defect for
φA = 0.4, H 0(cid:107) = 0.5 and J = 0. (b) Tubes without defects at J = 3. ( the bottom panel shows an enlarged side
view of the tip of a tube without defects ). (c) Tubes and buds with defects when J = 5. ( the bottom panel
shows an enlarged top view of the tip of a bud with defects )
F. Curvature differences leads to segregation
(a) Tubular membrane of uniform cross section with fields A and B in the mixed phase for H 0(cid:107),A, H 0(cid:107),B =
FIG. 8:
0.5, 0.5. (b) Activation of field B takes H 0(cid:107),B from 0.5 to 1.25. The difference in the spontaneous curvatures leads
to phase segregation and results in the constriction of tubes. Field concentrations are respectively φA = 0.4 and
φB = 0.6 while κ = 10kBT , κA(cid:107) = κB(cid:107) = 5kBT and AA = BB = 3kBT .
Another example of the curvature driven aggregation is demonstrated in Fig. 8. Shown in Fig. 8(a)
is a tubular membrane of uniform cross section, fully decorated by two different types of membrane
nematogens, labelled A and B. The tube is stable, in the mixed state, when the directional spontaneous
curvature of the in-plane fields, H 0(cid:107),A = H 0(cid:107),B = 0.5, are the same. If a source of activation increases the
spontaneous directional curvature of B to H0(cid:107),B = 1.25, analogous to activation of dynamin proteins
by hydrolysis of GTP, the fields demix. The regions containing field B constrict the tube further. The
equilibrium shape of the activated membrane is observed to have successive tubular regions of large
and small curvatures ( see Fig. 8(b)), similar to the striated patterns of dynamin tubes obtained on
treatment with GTPγs [55, 56]. For dynamin the molecular conformation and membrane tube diameter
is GTP dependent [56, 57]. Furthermore it is observed that the tube constriction involves a tube twisting,
suggesting a change in the helical angle ϕ [58]. In in vivo experiments, it has been demonstrated that
structurally similar F-BAR proteins can co-localize into the same membrane tubes [14] while differing
BAR proteins, like F-BAR and N-BAR, seggregate into separate membrane tubes with their respective
characteristic r and ϕ [41, 59]. Our analysis suggests that this recruitement of differing BAR-proteins
into separate domains is possibly driven by their directional spontaneous curvatures.
11
top viewDefectless tipside viewpair of +1/2 defects(a) J=0(b) J=3(c) J=5planetangentvN(v)c1vc2vt2vt1vvn(v)ϕvt2vt1v(a)(b)(c) Mixed phasedemixing and constrictionActivation of field B(a)(b)G. Sheets versus tubes
The effect of concentration, shown in Fig.(3), for surface coverages in the range 10% and 70% display
a series of complex shape deformations connected to different aggregate structures. The regime, where
inclusions stay separate, for the model parameters chosen here, appear at very low concentrations. The
figure illustrates that for a system where the direct interactions parameter J between the membrane
nematogens are weak the oligomers tend form larger rim-like formations, which stabilizes disc like struc-
tures of vesicles. The rims form the edges of the discs. As the concentration is increased part of the edge
turn tubular.
So for a range of concentrations the disc and the tubes coexist. The tubules get more pronounced and
the discs diminishes with the increase in concentration of membrane nematogens. Recent experiments
on the formation of tubular (or smooth) ER suggest that some membrane curvature active proteins,
reticulon protein and DP1[15], are highly enriched in the tubular ER [60] and the ER sheet edges[16].
Our results are thus in line with the idea that the concentration of these membrane nematogens are a
major determinant for the amount of ER in sheets or tubules[16].
IV. CONCLUSION
We have described the membrane curvature modifying properties of anisotropic protein inclusions,
like the BAR proteins, in terms of an in-plane nematic field. We have shown that the flexibility of
the membrane can promote aggregation and lateral domain formation of these membrane nematogens,
even in the absence of self interactions. These domains can facilitate shape changes of the membrane.
The equilibrium shapes obtained are strikingly similar to that seen in experiments involving curvature
modifying proteins. Prominent structures seen are tubes and discs and coexistence of them. Depending
on the preferred curvature of the nematogens, a protein lattice with helical nematic orientation around the
tube is seen. The properties of this liquid crystalline structure was further analyzed from a continuum
version of the model and the dependence of the tube radius and the orientation of the nematic with
respect to the tube axis was calculated. We also estimate the thermally induced fluctuation in these
quantities and show that they are comparable to what is seen in experiments. In addition we calculate
the persistence length of the nematogen induced tubes and show that it is in the range of experimentally
obtained values. This analysis provides the necessary basis to obtain estimates of model parameters from
experiments on coated membrane tubes. At present the available experimental data are very limited.
The present modeling provides additional support to the growing notion of the importance of local
curvature modulating proteins in membrane shape generation in biological cells. Compared to previous
modeling of the role of membrane proteins inducing directional membrane curvature we have taken into
account that membranes are not fully decorated, the in-plane interactions between nematogens and the
arbitrary membrane shapes with spherical topology. The current work focusses only on the membrane
interacting part of the protein. Electric charges in BAR-protein are mostly localized to its membrane
facing domain which in turn interact with anionic lipids and enables them to bind strongly to the
membrane. One natural extension of this model is include the electrostatic interactions through cytosol
between proteins moieties protruding out of the membrane.
We emphasize that the main aim of this work is to show that anisotropic curvature induced by the
inclusions can lead to aggregation and interesting shape changes. This is in contrast to the prevailing
assumption that explicit protein interactions are essential for aggregation and formation of protein lattices.
Though a quantitative comparison between the predictions of this model and experiments is not so easy,
the model does demonstrate the possibility of generating many biologically relevant shapes of the vesicle
by membrane mediated interactions alone.
12
V. ACKNOWLEDGEMENTS
The computational work is carried out at HPC facility at IIT-Madras.
[1] Frost A., V.M. Unger, and P. De Camilli, 2009, The BAR Domain Superfamily: Membrane-Molding Macro-
molecules. Cell 137, 191-196.
[2] Opali´nski (cid:32)L., Kiel J. A.K.W., C. Williams, M. Veenhuis, and I.J. van der Klei, 2011, Membrane curvature
during peroxisome fission requires Pex11. The EMBO Jour. 30, 5.
[3] Antonny B, 2006, Membrane deformations by protein coats. Curr. Opinion in Cell Biol. 18, 386.
[4] Sackmann E., R. Kotulla, and F-J. Heiszler, 1984, On the role of lipid-bilayer elasticity for the lipid-protein
interaction and the indirect protein-protein coupling. Canadian Jour. of Biochem. and Cell Biol. 62, 778
[5] Leibler S., 1986, Curvature instability in membranes. J. Phys. France 47, 507-516.
[6] Ramakrishnan N, P.B. Sunil Kumar, J.H. Ipsen, 2010, Monte Carlo simulations of fluid vesicles with in-plane
orientational ordering. Phys. Rev. E 81, 041922.
[7] Peter B.J., H.M. Kent, I.G. Mills,Y. Vallis, P.J.G. Butler, P.R. Evans and H.T. McMahon, 2004, BAR
Domains as Sensors of Membrane Curvature: The Amphiphysin BAR Structure. Science 303, 495-499.
[8] Weissenhorn W., 2005, Crystal structure of the Endophilin-A1 BAR Domain. J. Mol. Biol. 351, 653-661.
[9] McMohan H.T. and J. L. Gallop, 2005, Membrane curvature and mechanisms of dynamic cell membrane
remodelling. Nature 438, 590-596.
[10] Blood P.D. and G.A. Voth, 2006, Direct observation of Bin/amphiphysin/Rvs (BAR) domain-induced mem-
brane curvature by means of molecular dynamics simulations Proc. Natl. Acad. Sci. 103, 15068-15072
[11] Arkhipov A., Y. Ying, and K. Schulten, 2008, Four-scale description of membrane sculpting by BAR domains.
Biophys. J. 95, 2806-2821,
[12] Shnyrova A.V., V. A. Frolov and J. Zimmerberg, 2009, Domain-Driven Morphogenesis of Cellular Membranes.
Current Biology 19, R772-R780.
[13] Williams T.M. and M.P. Lisanti, 2004, The caveolin proteins. Genome Biology 5,214,1-8.
[14] Shimada A, H. Niwa, K. Tsujita, S. Suetsugu, K. Nitta, K. Hanawa-Suetsugu, R. Akasaka, Y. Nishino, M.
Toyama, L. Chen, Z.J. Liu, B.C. Wang , M. Yamamoto, T. Terada, A. Miyazawa, A. Tanaka, S. Sugano, M.
Shirouzu, K. Nagayama, T. Takenawa, S. Yokoyama, 2007, Curved EFC/F-BAR-domain dimers are joined
end to end into a filament for membrane invagination in endocytosis. Cell 129, 761-72.
[15] Voeltz G.K., W.A. Prinz, Y. Shibata, J.M. Rist and T.A. Rapoport, 2006, A class of membrane proteins
shaping the tubular endoplasmatic reticulum, Cell 124, 573-586.
[16] Shibata Y., T. Shemesh, W.A. Prinz, A.F. Palazzo, M.M. Kozlov and T.A. Rapoport, 2010, Mechanisms
Determining the Morphology of the Peripheral ER. Cell 143, 774-788.
[17] Farsaq K., N. Ringstad, K. Takei, S.R. Floyd, K. Rose, and P. De Camilli, 2001, Generation of high-curvature
membranes mediated by by direct endophilin bilayer interactions. J. Cell Biol. 155, 193-200.
[18] Gallop J. L., C.C. Jao, H.M. Kent, J.G. Butler, P.R. Evans, R. Langen and H.T. McMahon, 2006, Mechanism
of endophilin N-BAR domain mediated membrane curvature. The EMBO Journal 25, 2898-2910.
[19] Bouvrais H., K. J. Jensen, J. Brask, P. M´el´eard, T. Pott and J.H. Ipsen, 2008, Softening of POPC membranes
by Magainin. Biophysical Chemistry 137 7-12.
[20] Campelo F., McMahon H.T and M.M. Kozlov, 2008, The hydrophobic insertion mechanism of membrane
curvature generation by proteins Biophys. J. 95, 2325-2339
[21] Kurten R.C., A.D. Eddington , P. Chowdhury, R.D. Smith, A.D. Davidson and B.B. Shank, 2001, Self-
assembly and binding of a sorting nexin to sorting endosomes. J. Cell Science 114, 1743-1756.
[22] Mayor S., J.F. Presley and F.R. Maxfield, 1993, Sorting of Membrane Components from Endosomes and
Subsequent Recycling to the Cell Surface Occurs by a Bulk Flow Process. J. Cell. Biol. 121,1257-1269.
[23] Shepard, K.A. and M.P. Yaffe, 1999, The Yeast Dynamin-like Protein, Mgm1p, Functions on the Mitochon-
drial Outer Membrane to Mediate Mitochondrial Inheritance. J. Cell. Biol. 144, 711-720.
[24] Razzaq A., I.M. Robinson, H.T. McMahon, J.N. Skepper, Y. Su1, A.C. Zelhof, A.P. Jackson, N.J. Gay
and C.J. O'Kane, 2001, Amphiphysin is necessary for organization of the excitation contraction coupling
machinery of muscles, but not for synaptic vesicle endocytosis in Drosophila. Genes & Dev. 15, 2967-2979
[25] Sweitzer S.M. and J.E. Hinshaw, 1998, Dynamins Undergoes a GTP-Dependent Conformational Change
Causing Vesiculation. Cell 98, 1021-1029.
[26] Low H.H., C. Sachse, A.A. Amos, and J. Lowe, 2009, Structure of a Bacterial Dynamin-like Protein Lipid
Tube Provides a Mechanism For Assembly and Membrane Curving. Cell 139, 1342-1352.
[27] Lebbink M.N., N. Jim´enez, K. Vocking, L.H. Hekking, A.J. Verkleij and J.A. Post. 2010. Spiral Coating of
the Endothelial Caveolar Membranes as Revealed by Electron Tomography and Template Matching. Traffic
11, 138-150.
[28] Sens P. and M. S. Turnery, 2004, Theoretical Model for the Formation of Caveolae and Similar Membrane
13
Invaginations. Biophys. J. 86, 2049-2057.
[29] Fournier J. -B., 1996, Nontopological saddle-splay and curvature instabilities from anisotropic membrane
inclusions. Phys. Rev. Lett., 76(23), 4436.
[30] Fournier J. -B. and P. Galatola, 1998, Bilayer Membranes with 2D-Nematic Order of the Surfactant Polar
Heads. Braz. Jour. of Phys., 28, 329.
[31] Kraij-Iglic V., V. Heinrich, S. Svetina, S. and B. Zeks, 1999, Free energy of closed membrane with anisotropic
inclusions. Eur. Phys. J. B 10, 5-8.
[32] Dommersnes P. G. and J.-B. Fournier, 1999, N-body study of anisotropic membrane inclusions: Membrane
mediated interactions and ordered aggregation Eur. Phys. J. B 12, 9-12.
[33] Iglic A., T. Slivnik, and V. Kralj-Iglic, 2007, Elastic properties of biological membranes influenced by attached
proteins. J. Biomech. 40, 2492-2500.
[34] Helfrich W., 1973, Elastic properties of lipid bilayers: theory and possible experiments. Naturforsch Z. 28,
693-703.
[35] Onsager L., 1949, The effects of shape on the interaction of colloidal particles. Annals of the New York
Academy of Sciences 51, 627-659.
[36] Park J.-M and T. C Lubensky, 1996, Interactions between membrane Inclusions on Fluctuating Membranes.
Journal de Physique I 6, 1217.
[37] Kim K. S., J. Neu, and G. Oster, 1998, Curvature-mediated interactions between membrane proteins. Biophys.
J. 75, 2274.
[38] Chou T., K. S. Kim and G. Oster, 2001, Statistical thermodynamics of membrane bending-mediated protein-
protein attractions. Biophys. J. 80, 1075.
[39] Lewandowski, E. P., J. A. Bernate, P. C. Searson, P C and K. J. Stebe, 2008, Rotation and Alignment of
Anisotropic Particles on Nonplanar Interfaces, Langmuir, 24, 9302.
[40] Lewandowski, E. P., J. A. Bernate, A. Tseng, P. C. Searson, P C and K. J. Stebe, 2009, Oriented assembly
of anisotropic particles by capillary interactions, Soft Matter, 5, 886.
[41] Frost A., R. Perera, A. Roux, K. Spasov, O. Destaing, E.H. Engelmann, P. De Camilli, and A, Unger, 2008,
Structural Basis of Membrane Invaginations by F-BAR Domains. Cell 132, 807-817.
[42] Peliti L. and J. Prost, 1989, J.Phys. France 50, 1557-1571.
[43] Frank J. R. and M. Kardar, 2008, Defects in nematic membranes can buckle into pseudospheres, Phys. Rev.
E 77, 041705.
[44] Kumar P. B. S. and Madan Rao, 1998, Shape Instabilities in the Dynamics of a Two-Component Fluid
Membrane, Phys. Rev. Lett. 80, 2489.
[45] Kumar P. B. S., Gerhard Gompper and Reinhard Lipowsky, 2001, Budding Dynamics of Multicomponent
Membranes, Phys. Rev. Lett. 86, 3911.
[46] Henriksen J.R., T.L. Andresen, L. Feldborg, L. Duelund, and J.H. Ipsen, 2010, Understanding Detergent
Effects on Lipid Membranes: A Model Study. Biophy. J. 98, 2199-2205.
[47] Gerbeaud C, 1998, Effect de l'insertion de prot´eines et de peptides membranaires sur les propri´et´es m´ecaniques
et les changements morphologiques de v´esicules g´eantes. Ph.d. thesis L'Universte Bordeaux I.
[48] Gouttenoire J., P. Roingeard, F. Penin, and D. Moradpour1, 2010, Amphipathic α-Helix AH2 Is a Major
Determinant for the Oligomerization of Hepatitis C Virus Nonstructural Protein 4B. Journal of Virology 84,
12529-12537.
[49] Bonifacino J.S. and J.H. Hurley, 2008, Retromer, Curr. Opinion in Cell Biol. 20, 427-436
[50] Sweitzer S.M. and J.E. Hinshaw, 1998, Dynamin undergoes a GTP-dependent conformational change causing
vesiculation. Cell 93, 1021.
[51] Sorre B, A. Callan-Jones, J. Manzi, B. Goud, J. Prost, P. Bassereau, A. Roux, 2012, Nature of curvature
coupling of amphiphysin with membranes depends on its bound density. Proc. Natl. Acad. Sciences 109,
173-8.
[52] Lubensky T.C. and J. Prost, 1992, Orientational order and vesicle shape. J. Phys. France II 2, 371-382.
[53] Vitelli V. and D.R. Nelson, 2004, Defect generation and deconfinement on corrugated topographies. Phys.
Rev. E 70, 051105.
[54] Ramakrishnan N., J. H. Ipsen, P. B. Sunil Kumar, 2012, Role of disclinations in determining the morphology
of deformable fluid interfaces. Soft Matter 8, 3058-3061.
[55] Roux, A., G. Koster, M. Lenz, B. Sorre, J.-B. Manneville, P. Nassoy and P. Bassereau, 2010, Membrane
curvature controls dynamin polymerization. Proc. Natl. Acad. Sciences 107, 4141-4146.
[56] Daninoa D., K-H. Moonb and J.E. Hinshaw, 2004, Rapid constriction of lipid bilayers by the mechanochem-
icalenzyme dynamin. Journal of Structural Biol. 147, 259-267.
[57] Zhang P. and J.E. Hinshaw, 2001, Three-dimensional reconstruction of dynamin in the constricted state.
Nature Cell Biology 3, 922-926.
[58] Roux A., K. Uyhazi, A. Frost, and P. De Camilli, 2006, GTP-dependent twisting of dynamin implicates
constriction and tension in membrane fission. Nature 44, 528-531.
[59] Itoh T. and P. De Camilli, 2006, BAR, F-BAR(EFC), and ENTH/ANTH domains in the regulation of
membrane cytosol interfaces and membrane curvature. Biochim. Biophys. Acta 1761, 897-912
14
[60] Shibata Y., C. Voss, J.M. Rist, J. Hu, T.A. Rapoport, W.A. Prinz and G.K. Voeltz, 2008, The reticulon
and DP1/Yop1p proteins form immobile oligomers in the tubular endoplasmic reticulum. J. Biol. Chem. 283,
18892-904.
15
|
1007.2432 | 1 | 1007 | 2010-07-14T20:50:44 | Chiral Interactions of Histidine in a Hydrated Vermiculite | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.BM"
] | Recent work suggests a link between chiral asymmetry in the amino acid iso-valine extracted from the Murchison meteorite and the extent of hydrous alteration. We present the results of neutron scattering experiments on an exchanged, 1-dimensionally ordered n-propyl ammonium vermiculite clay. The vermiculite gel has a (001) d-spacing of order 5nm at the temperature and concentration of the experiments and the d-spacing responds sensitively to changes in concentration, temperature and electronic environment. The data show that isothermal addition of D-histidine or L-histidine solutions produces shifts in the d-spacing that are different for each enantiomer. This chiral specificity is of interest for the question of whether clays could have played an important role in the origin of biohomochirality. | physics.bio-ph | physics | Chiral Interactions of Histidine in a Hydrated Vermiculite Clay
Donald G. Fraser,1* H. Christopher Greenwell,2 Neal T. Skipper,3 Martin V.
Smalley,4 Michael A. W ilkinson,3,5 Bruno Demé,5 R. K. Heenan.6
1Department of Earth Sciences, University of Oxford, Parks Road, Oxford OX1 3PR, UK
2Department of Chemistry, Durham University, South Road, Durham DH1 3LE, UK
3Department of Physics & Astronomy, University College London, Gower Street, London
WC1E 6BT, UK
4Department of Physics, University of York, Heslington, York YO10 5DD, UK
5Institut Laue Langevin, F-38042 Grenoble, Cedex 9, France
6ISIS Facility, R3 1-22, Rutherford Appleton Lab, HSIC, Didcot, OX11 0QX, UK
Email: [email protected]
Graphical Contents Entry
We report chiral differences in the effects of D- and L-histidine on the interlayer
spacing of vermiculite clays that are relevant in assessing the role of clays in the
origin of bio-homochirality.
1
Abstract
Recent work suggests a link between chiral asymmetry in the amino acid iso -valine
extracted from the Murchison meteorite and the extent of hydrous alteration.
We present the results of neutron scattering experiments on an exchanged, 1 -
dimensionally ordered n-propyl ammonium vermiculite clay. The vermiculite gel has
a (001) d-spacing of order 5nm at the temperature and concentration of the
experiments and the d-spacing responds sensitively to changes in concentration,
temperature and electronic environment. The data show that isothermal addition of D-
histidine or L-histidine solutions produces shifts in the d-spacing that are different for
each enantiomer. This chiral specificity is of interest for the question of whether clays
could have played an important role in the origin of biohomochirality.
Introduction
Understanding the interaction of biomolecules such as amino acids and simple sugars
with hydrated mineral surfaces is key to unraveling the origins of life on Earth1. The
chirality of some crystalline mineral phases, such as quartz and calcite, has long been
proposed as a possible mechanism for selectively determining the production of
particular prebiotic enantiomers2. However, neither calcite nor quartz has been shown
to have any heterogeneous catalytic function in aiding biomolecule polymerization.
Fundamental questions arise as to how the first organic molecules arrived on Earth
and were built up into complex molecules through prebiotic chemistry, whilst being
protected from the harsh environment present at the time. Though the view that life
arose through an “RNA World” is widely supported3-4, there is also an alternative
hypothesis that proteins originated first on the early Earth5. Lambert has recently
reviewed the organization and polymerization of amino acids at mineral surfaces6.
Clay minerals have been shown to be efficient at concentrating amino acids from
solution7-9, especially those with charged side groups, such as lysine or histidine.
Clays have also been shown to catalyse peptide bond formation in their own right10-12,
and when used in conjunction with salt induced peptide formation (SIPF)
chemistries13-15.
Clay minerals in carbonaceous meteorites have also been shown to be of potential
importance for protecting organic molecules from oxidizing conditions16. At least 70
chiral amino acids have been identified in meteorites. These were believed to show
almost no chiral selectivity, being close to 50:50 racemic mixtures17, but recently an
important chiral asymmetry in iso-valine extracted from the Murchison meteorite has
been reported18. Laboratory gas condensation experiments also generate amino acids,
though, again, these show almost no chiral selectivity19. In contrast, amino acids in
proteins and enzymes in organisms are usually 100% L-enantio-specific.
Prebiotic chemistry must address the question of homochirality and the emergence
of only L-amino acids in virtually all living organisms. Goldberg showed, once
seeded with one enantiomer, clay surfaces enhance selectivity for that enantiomer
2
during crystallization cycles20. Recent experiments have pointed towards chirally
selective adsorption and reactivity in clay mineral suspensions21-22. Anionic clays
have also been shown to have kinetic chiral selectivity for histidine23. The anionic
clays have much higher charge density and packing than for the cationic systems. Yu
et al. used density functional theory computer simulations to examine the energy of L-
and D-alanine on the surface of nontronite clay, finding a preference by some 25
kJ/mol for the L- form23. Adsorption of amino acids at clay mineral surfaces is also of
interest in a range of natural phenomena24-26. Industrially, amino acid chiral selectivity
and racemization stabilization effects have been observed in anionic clays (layered
double hydroxides)27.
In many adsorption studies, structural analysis of the clay-amino acid system is
often conducted post reaction, whereas there are no high-resolution in situ structural
studies of fully hydrated systems. In the work presented here, we studied the addition
of D- and L-histidine to n-propylammonium vermiculite gels (with layer spacing ~50
Å). The alkylammonium vermiculite gels have been established as a model system for
studying clay swelling and colloid stability28. The swelling is nearly perfectly one-
dimensional, taking place perpendicular to the plane of the silicate layers, and leads to
the formation of gels with d-values between 40 and 90 Å along the swelling axis.
Even for the largest expansions, this d-value remains sufficiently well defined to be
measured by small-angle neutron diffraction (SANS).
Experimental
composition
Crystals
dry
of
vermiculite
Eucatex
natural
of
Si6.13Mg5.44Al1.65Fe0.50Ti0.13Ca0.13Cr0.01K0.01O20(OH)4Na1.29 were
exchanged
in
CH3CH2CH2NH2·HCl (Sigma-Aldrich 242543, >99%) solution over a period of
several years. Around two weeks before the experiments crystals of approximate
dimensions 10 × 10 × 2 mm were placed in stock solution of propylammonium
chloride in D2O at the required concentration, and were equilibrated for the neutron
scattering experiments with regular exchange of supernatant solution. Matched
solutions
containing D-histidine
(Puriss® 53321, ≥99%) or L-histidine
(ReagentPlus® H8000, ≥99% both analysed and certificated by TLC) were made by
addition of reagent to D2O or identical stock solution of propylammonium chloride.
Individual equilibrated gel crystals were selected from solution and placed in sample
containers. The samples were then bathed in solution as detailed below. The swollen
gels are extremely soft and delicate, and all handling was conducted under solution.
Previous studies of these gels have established the relationship between d-spacing and
concentration in propylammonium chloride solution28. In the region of most interest,
around 0.35M, the equilibrium d-spacing increases by around 1 Å for a 0.01M
decrease in concentration. The changes we report here are therefore significant when
compared with the reliability of the solution concentrations.
Preliminary experiments were conducted on the LOQ instrument at the ISIS Pulsed
Neutron source, Rutherford Appleton Laboratory. We equilibrated stocks of gels in
0.1M, 0.2M and 0.35M propylammonium chloride in D2O. At these concentrations
the d-spacings are approximately 110, 75 and 50 Å respectively. As far as possible
gels were selected from the stock as matched pairs, and when both showed a clear
3
(001) Bragg peak we then exposed one to L- and one to D-histidine. In this way we
studied 6 pairs in total; 1 at 0.1M, 2 at 0.2M and 3 at 0.36M.
Samples were placed flat on the bottom of sealed 10 × 10 mm Hellma QS cells and
were bathed in 3cm3 of the supernatant solution of the required concentration. The
sample height and alignment was adjusted individually by laser beam, and the small
angle neutron scattering was measured in the Q range 0.006 - 0.24 Å-1. After this
initial characterisation the samples were exposed to amino acid by the addition of
0.5cm3 of 0.18M L- or D-histidine in D2O. After approximately 1 hour the neutron
scattering was measured again. Temperature was maintained at 30 ±1C throughout
by a water bath in thermal contact with the sample block. Cycles through a
temperature range were conducted to confirm the crystal-gel transition.
Scattering intensity was corrected for background from the empty container and
normalised to an absolute scale by reference to a water standard and a beam
transmission measurement. Due to the sample orientation the Bragg scattering from
the (00l) reflections was recorded in the top and bottom quadrants of the area detector.
The data were binned in Q with a constant bin-width of 0.001 Å-1. The scattered
intensities were fitted with the program Fish [Heenan, R. K. Fish data analysis
program, RAL-Report-89-129, Rutherford Appleton Laboratory (UK) 1989] using
Lorentzian Bragg profiles with a form factor derived for a distribution of oriented
sheets29-30. Example fits are shown in Figure 1.
In all cases the d-spacing was increased at constant concentration by exposure to
histidine. If we express this expansion in terms of the fractional change
then the
average shift is 1.117 ±0.020 for L-histidine and 1.102 ±0.020 for D-histidine. The
average expansion was therefore 1.5% greater for L-histidine than D-histidine, but
tantalisingly, this is on the limit of the statistical precision of the experiments.
Figure 1: Example of neutron scattering intensity I (arbitrary units) obtained from the
instrument LOQ for the sample exposed to 0.1M propyl ammonium chloride before
exposure to L-histidine. Data (+) and fit (line). Inset shows intensity I on a log10 scale.
4
Following these initial studies on the LOQ instrument at ISIS we exploited the
increased Q-range of station D16 at the Institut Laue-Langevin, Grenoble, to measure
1st and higher order (00l) peaks in both the crystalline and gel phases of
propylammonium ion substituted vermiculite clay. The instrument was operated at a
constant wavelength of 4.728 Å with a pixelated detector giving a Q-range of 0.04 –
0.5 Å-1 with a bin-width of 0.0025 Å-1. The detector efficiency was normalized by
reference to a water standard. Samples of gel phase propylammonium vermiculite
were prepared via the method of Williams et al.31-32, by soaking crystals in a
deuterated solution of 0.34M propylammonium chloride. We first measured the
structure of the gels in the absence of amino acid as a control. Clay gels were
equilibrated in 0.3387M propylammonium chloride in D2O and a high-quality sample
was cleaved into two halves (Figure 2). Each half was then placed in a QS cell with 4
× 10 mm sample space under supernatant solution and was aligned using a laser beam
and then via a rocking curve centred on the (001) Bragg reflection. Temperature was
maintained at 25.5 ±0.5C via a thermal bath connected to a neutron-transparent
thermal jacket.
Figure 2. Vermiculite clay gel in solution for measurement on D16.
After measurement on the samples under pure 0.3387M propylammonium chloride in
D2O, the solution was replaced with 0.3387m propylammonium chloride containing
either 0.05M L- or D-histidine in D2O. After 5 full exchanges of solution over the
samples, conducted over a period of around 6 hours during which the neutron
scattering profile was monitored, steady state equilibrium was achieved. The final
profile was then measured.
The normalized intensities were fitted as before using the program Fish using
Lorentzian Bragg profiles with a form factor derived for a distribution of oriented
sheets. A typical fit is shown in Figure 3.
5
Figure 3: Example of neutron scattering intensity I (arbitrary units) obtained from the
instrument D16 for the sample exposed to 0.3387M propyl ammonium chloride and
0.05M L-histidine. Data (+) and fit (line). Inset shows intensity I on a log10 scale.
Discussion
It is known that the phase transition between these two phases occurs at around
310K28. Our hypothesis was that the thermodynamics of the phase diagram and the
interlayer d-spacings of the clay, would be sensitive to the adsorption of histidine
enantiomers onto the clay surface, and that we should therefore be able detect their
uptake/location and any preferential chiral interaction. Figure 4 shows data obtained
for the two nearly identical halves of untreated gel measured first in pure 0.34M
propylammonium solution. These can be compared in the same Figure with the data
for the same vermiculite samples measured after 5 exchanges in L- and D-histidine
respectively.
The peak in all cases arises from the (001) Bragg reflection – and therefore directly
gives the d-spacing of the sample. Increase in ordinary solute concentration would
cayuse the d-spacing to contract. The data in Figure 4 show clearly that the amino
acids act “anti-osmotically” so that addition of amino acid, and hence increase in the
overall solute concentration, causes the d-spacing to widen relative to that of the pure
control clay system. This suggests that histidine adsorbs to the clay surface, screening
the Coulombic charge. In addition, there is evidence of a chiral shift: L-histidine
increases the d-spacing relative to D-histidine. The value for the control was 48.78Å;
D-histidine shifted from 49.06 to 50.33Å; L-histidine shifted from 48.76 to 51.11Å.
The ~1.5% difference between L- and D- is significant in terms of the accuracy of the
instrument, and is also consistent with our previous experiments on LOQ.
To provide further information on the nature of the amino acid adsorption, a separate
experiment was conducted in which a previously L-histidine-equilibrated sample was
exposed to D-histidine. In this case there was no significant reverse shift in the d-
spacing on going from L- to D- over the timescale of the experiment (~ 6 hours). This
lack of short term reversibility lends support to the notion that the amino acids are
bound directly and strongly to the clay surface.
6
Figure 4. Shift in d-spacing of histidine intercalated propyl-ammonium vermiculite
clay gels and control systems. L-histidine and control: filled and empty circles. D-
histidine and control: filled and empty triangles.
The behaviour of the histidine molecules in the inter-layer region of the hydrated
amino acid clay system was modeled by Monte-Carlo and molecular dynamics
simulations and a schematic of the interlayer region of the hydrated clay is shown in
Fig.5. The Monte Carlo algorithm Monte was used to generate the initial coordinates
for a hydrated propylammonium vermiculite with phenol in the interlayer. Forcefield
parameters were the OPLS. The equilibrated system was taken and the phenol
molecules converted to histidine. 40 ps of molecular dynamic simulation was then
run using the Dreiding forcefield, with the charges previously assigned, with the
Forcite code in the Accelrys Inc Materials Studio. It is apparent that a large
proportion of the histidine molecules are indeed adsorbed at the mineral surfaces as
suggested above.
Figure 5. Schematic based on molecular simulations showing interlayer arrangement
of histidine (green coloured) molecules in a propylammonium (blue molecules)
vermiculite clay (color: Si = orange; Al = magenta; Mg = green; O = red; H = white).
7
To try to determine whether the effects described above were due to preferential
uptake of one isomer over the other, chiral high performance liquid chromatography
(HPLC) was undertaken. Experiments were carried out in 8 ml screwtop glass vials.
A 250ml stock solution of 0.40mg/ml DL histidine and 0.338M PrNH3Cl was
prepared. Raw Eucatex vermiculite, which had been stored under distilled water, was
dried at 80 C for 24 hours. Sodium vermiculite was exchanged twice over 24 hours
with 0.388M propylamine solution, and dried at 140 C for 1 hour. Different weights
of clay were then added to a constant amount of histidine solution, so that the change
in histidine concentration could be found. The raw vermiculite samples were left to
equilibrate over 24 hours, whilst the propylamine vermiculite equilibrated over 1.5
hours.
Analysis method A Perkin-Elmer Peltier plate HPLC with autosampler, in the
Analytical Laboratory, Department of Chemistry, Durham University, was used for
all analyses. A Crownpak CR(+) chiral HPLC column (D form eluted first) was used
throughout. Stock 20% perchloric acid diluted to give pH 1 was used as an eluent,
with the histidine underivatised. A temperature of 10 C and a flow rate of 0.1
ml.min-1 was found to give good separation and could be reliably maintained. An UV
detector, 200 nm, was used. The retention times were 4.73 min and 6.27 min for the
D- and L-histidine, respectively.
A set of standards was made up at the experimental concentration of 0.388M and with
the change in enantiomeric excess (ee) set at 5.00%, 1.00%, 0.50% and 0.05%. For a
mixture of isomers the relative area under each peak, expressed as a percentage of the
total of the two isomer peaks, indicates the enantiomeric excess. At 5% and 1% ee the
steps were reasonably well resolved. The relative high values of the area measured
(in V. s-1) lead to a spread in the results at the lower ee of 0.5% and 0.05%.
However, when these are converted to percentage ee, a clearer trend was defined.
Even at 0.5% ee, the observed changes seem valid. In summary, it would seem that
the instrument and method used were capable of discerning, within reasonable error,
increments down to 0.05%, though the absolute value may be slightly different to the
standard.
To investigate the effects of adsorption on the vermiculite clay varying weights of
clay were added to the racemic mix of D/L histidine and the supernatant measured for
preferential adsorption of amino acid. In addition to testing propylammonium
exchanged clay as used in the neutron scattering experiments, the sodium clay and
propylammonium solution were tested, and the sodium clay alone was tested to
determine any effect the propyl ammonium clay / solution may have had.
Within the error present within the data, and allowing for the slight non-equivalence
in the baseline solutions with no clay present, it is not possible to discern an ee in
favour of either enantiomer. The size of the variation in the data was +/- 0.3 % ee .
Thus, within the experimental error, we could find no evidence for preferential
intercalation of either enantiomer. This suggests that the observed shift in d-spacing
is purely due to subtle electrostatic screening effects between the two isomers.
In conclusion, we demonstrated that the effect of different enantiomers of an amino
acid may alter physical and structural properties of clay systems, at low amino acid
concentrations and over a long length-range. This considerably extends the previous
studies of chiral selectivity by clays, which usually require high packing densities and
8
low hydration states for differences to become apparent. HPLC analysis was unable to
discern any compositional differences between the different systems, indicating that
the observed effect is due to structural differences within the interlayer.
4.
3.
5.
6.
7.
8.
Acknowledgements. We are grateful to ISIS and ILL for neutron beamtime. We
thank Dr A. Congreve, Department of Chemistry, Durham University, for assistance
with HPLC measurements. DGF gratefully acknowledges financial support from
Worcester College, Oxford.
References
1.
2.
D. G. Fraser, Cell Origin Life Ext, 2004, 7, 149-152.
R. M. Hazen, T. R. Filley and G. A. Goodfriend, P Natl Acad Sci USA, 2001, 98,
5487-5490.
R. F. Gesteland, T. Cech and J. F. Atkins, The RNA world : the nature of modern
RNA suggests a prebiotic RNA world, 3rd edn., Cold Spring Harbor Laboratory Press,
Cold Spring Harbor, N.Y., 2006.
R. F. Gesteland and J. F. Atkins, The RNA world : the nature of modern RNA
suggests a prebiotic RNA world, Cold Spring Harbor Laboratory Press, Cold Spring
Harbor, NY, 1993.
K. Plankensteiner, H. Reiner and B. M. Rode, Curr Org Chem, 2005, 9, 1107-1114.
J. F. Lambert, Origins Life Evol B, 2008, 38, 211-242.
A. Parbhakar, J. Cuadros, M. A. Sephton, W. Dubbin, B. J. Coles and D. Weiss,
Colloid Surface A, 2007, 307, 142-149.
L. O. B. Benetoli, C. M. D. de Souza, K. L. da Silva, I. G. D. Souza, H. de Santana,
A. Paesano, A. C. S. da Costa, C. T. B. V. Zaia and D. A. M. Zaia, Origins Life Evol
B, 2007, 37, 479-493.
J. Ikhsan, B. B. Johnson, J. D. Wells and M. J. Angove, J Colloid Interf Sci, 2004,
273, 1-5.
J. Bujdak, H. LeSon and B. M. Rode, J Inorg Biochem, 1996, 63, 119-124.
J. Bujdak and B. M. Rode, J Mol Catal a-Chem, 1999, 144, 129-136.
J. Cuadros, L. Aldega, J. Vetterlein, K. Drickamer and W. Dubbin, J Colloid Interf
Sci, 2009, 333, 78-84.
H. Le Son, Y. Suwannachot, J. Bujdak and B. M. Rode, Inorg Chim Acta, 1998, 272,
89-94.
B. M. Rode, H. L. Son, Y. Suwannachot and J. Bujdak, Origins of Life and Evolution
of the Biosphere, 1999, 29, 273-286.
F. Li, D. Fitz, D. G. Fraser and B. M. Rode, Amino Acids, 2010, 38, 287-294.
L. A. J. Garvie and P. R. Buseck, Meteorit Planet Sci, 2007, 42, 2111-2117.
O. Botta, D. P. Glavin, G. Kminek and J. L. Bada, Origins Life Evol B, 2002, 32, 143-
163.
D. P. Glavin and J. P. Dworkin, Meteorit Planet Sci, 2009, 44, A78-A78.
G. M. M. Caro, U. J. Meierhenrich, W. A. Schutte, B. Barbier, A. A. Segovia, H.
Rosenbauer, W. H. P. Thiemann, A. Brack and J. M. Greenberg, Nature, 2002, 416,
403-406.
S. I. Goldberg, Origins of Life and Evolution of the Biosphere, 2007, 37, 55-60.
B. Siffert and A. Naidja, Clay Miner, 1992, 27, 109-118.
T. Ikeda, H. Amoh and T. Yasunaga, J Am Chem Soc, 1984, 106, 5772-5775.
C. H. Yu, S. Q. Newton, D. M. Miller, B. J. Teppen and L. Schafer, Struct Chem,
2001, 12, 393-398.
X. C. Wang and C. Lee, Mar Chem, 1993, 44, 1-23.
24.
N. Lahajnar, M. G. Wiesner and B. Gaye, Deep-Sea Res Pt I, 2007, 54, 2120-2144.
25.
J. I. Hedges and P. E. Hare, Geochim Cosmochim Ac, 1987, 51, 255-259.
26.
27. M. Wei, Q. Yuan, D. G. Evans, Z. Q. Wang and X. Duan, J Mater Chem, 2005, 15,
1197-1203.
20.
21.
22.
23.
10.
11.
12.
9.
13.
14.
15.
16.
17.
18.
19.
9
29.
28. M. Smalley, Clay swelling and colloid stability, CRC/Taylor & Francis, Boca Raton,
Fla. ; London, 2006.
N. T. Skipper, A. K. Soper and J. D. C. McConnell, J Chem Phys, 1991, 94, 5751-
5760.
30. M. Kotlarchyk and S. M. Ritzau, Journal of Applied Crystallography, 1991, 24, 753-
758.
G. D. Williams, N. T. Skipper and M. V. Smalley, Physica B, 1997, 234, 375-376.
G. D. Williams, A. K. Soper, N. T. Skipper and M. V. Smalley, J Phys Chem B,
1998, 102, 8945-8949.
31.
32.
10
|
1606.05973 | 1 | 1606 | 2016-06-20T05:13:28 | A New Parallel Algorithm for Two-Pass Connected Component Labeling | [
"cs.DS",
"cs.CV",
"cs.PF"
] | Connected Component Labeling (CCL) is an important step in pattern recognition and image processing. It assigns labels to the pixels such that adjacent pixels sharing the same features are assigned the same label. Typically, CCL requires several passes over the data. We focus on two-pass technique where each pixel is given a provisional label in the first pass whereas an actual label is assigned in the second pass.
We present a scalable parallel two-pass CCL algorithm, called PAREMSP, which employs a scan strategy and the best union-find technique called REMSP, which uses REM's algorithm for storing label equivalence information of pixels in a 2-D image. In the first pass, we divide the image among threads and each thread runs the scan phase along with REMSP simultaneously. In the second phase, we assign the final labels to the pixels. As REMSP is easily parallelizable, we use the parallel version of REMSP for merging the pixels on the boundary. Our experiments show the scalability of PAREMSP achieving speedups up to $20.1$ using $24$ cores on shared memory architecture using OpenMP for an image of size $465.20$ MB. We find that our proposed parallel algorithm achieves linear scaling for a large resolution fixed problem size as the number of processing elements are increased. Additionally, the parallel algorithm does not make use of any hardware specific routines, and thus is highly portable. | cs.DS | cs | A New Parallel Algorithm for Two-Pass Connected
Component Labeling
Department of Electrical Engineering & Computer Science, Northwestern University, Evanston, IL 60208, USA
Siddharth Gupta, Diana Palsetia, Md. Mostofa Ali Patwary, Ankit Agrawal, Alok Choudhary
[email protected], {drp925, mpatwary, ankitag, choudhar}@eecs.northwestern.edu
1
6
1
0
2
n
u
J
0
2
]
S
D
.
s
c
[
1
v
3
7
9
5
0
.
6
0
6
1
:
v
i
X
r
a
Abstract-Connected Component Labeling (CCL)
is an
important step in pattern recognition and image processing. It
assigns labels to the pixels such that adjacent pixels sharing
the same features are assigned the same label. Typically, CCL
requires several passes over the data. We focus on two-pass
technique where each pixel is given a provisional label in the
first pass whereas an actual label is assigned in the second pass.
We present a scalable parallel two-pass CCL algorithm, called
PAREMSP, which employs a scan strategy and the best union-find
technique called REMSP, which uses REM's algorithm for storing
label equivalence information of pixels in a 2-D image. In the
first pass, we divide the image among threads and each thread
runs the scan phase along with REMSP simultaneously. In the
second phase, we assign the final labels to the pixels. As REMSP
is easily parallelizable, we use the parallel version of REMSP for
merging the pixels on the boundary. Our experiments show the
scalability of PAREMSP achieving speedups up to 20.1 using 24
cores on shared memory architecture using OpenMP for an image
of size 465.20 MB. We find that our proposed parallel algorithm
achieves linear scaling for a large resolution fixed problem size
as the number of processing elements are increased. Additionally,
the parallel algorithm does not make use of any hardware specific
routines, and thus is highly portable.
I. Introduction
One of
the most
fundamental operations in pattern
recognition is the labeling of connected components in a
binary image. Connected component labeling (CCL) is a
procedure for assigning a unique label to each object (or
a connected component) in an image. Because these labels
are key for other analytical procedures, connected component
labeling is an indispensable part of most applications in
pattern recognition and computer vision, such as fingerprint
identification, character recognition, automated inspection,
target recognition, face identification, medical image analysis,
and computer-aided diagnosis. In many cases, it is also one of
the most time-consuming tasks among other pattern-recognition
algorithms [1]. Therefore, connected component
labeling
continues to be an active area of research [2]–[9].
There exist many algorithms for computing connected
components in a given image. These algorithms are categorized
into mainly four groups [10] : 1) repeated pass algorithms
[11], [12], 2) two-pass algorithms [13]–[21] 3) Algorithms
with hierarchical tree equivalent representations of the data
[22]–[29], 4) parallel algorithms [30]–[35]. The repeated
pass algorithms perform repeated passes over an image in
forward and backward raster directions alternately to propagate
the label equivalences until no labels change. In two-pass
algorithms, during the first pass, provisional labels are assigned
to connected components; the label equivalences are stored in
a one-dimensional or a two-dimensional table array. After the
first pass, the label equivalences are resolved by some search.
This step is often performed by using a search algorithm
such as the union-find algorithm. The results of resolving
are generally stored in a one-dimensional table. During the
second pass, the provisional labels are replaced by the smallest
equivalent label using the table. Since the algorithm traverses
image twice, these algorithms are called two-pass algorithms.
In algorithms that employ hierarchical tree structures i.e.,
n-ary tree such as binary-tree, quad-tree, octree, etc., the label
equivalences are resolved by using a search algorithm such
as the union-find algorithm. Lastly, the parallel algorithms
have been developed for parallel machine models such as a
mesh connected massively parallel processor. However all these
algorithms share one common step, known as scanning step in
which provisional label is given to each of the pixel depending
on its neighbors.
In this paper we focus on two-pass CCL algorithms. The
algorithm in [36] and [37] are two developed techniques for
two-pass Connected Component Labeling. The algorithm in
[36], which we refer to as CCLLRPC, uses a decision tree to
assign provisional labels and an array-based union-find data
structure to store label equivalence information. However, the
technique employed for union-find, Link by Rank and Path
Compression is not the best technique available [38]. The
algorithm in [37], which we refer to as ARUN, employs a
special scan order over the data and three linear arrays instead
of the conventional union-find data structure. There exists
a parallel implementation of ARUN on TILE64 many core
platform [39]. According to the experimental results given in
[39], the parallel implementation is able to achieve a speedup
of 10 on 32 processor units. This implementation is also
not portable due to its implementation for specific hardware
architecture.
We propose two two-pass algorithms for labeling the
connected components, AREMSP and CCLREMSP, which
are based on REM's union-find algorithm REMSP [40], [41]
and the scan strategy of ARUN and CCLLRPC algorithms.
Since REM's union-find is an algorithm which implements
immediate parent check test and compression technique called
Splicing [40], [41], our proposed sequential two-pass algorithm
AREMSP is 39% faster than CCLLRPC and 4% faster than
ARUN. Another advantage of using REM's union-find approach
is that its parallel implementation is shown to scale better
with increasing number of processor [38]. Parallel REM's
TABLE I: Abbreviations used in the paper and their brief description
Abbreviation
CCL
ARUN
REMSP
AREMSP
PAREMSP
CCLLRPC
CCLREMSP
Description
Connected Component Labeling
CCL algorithm suggested by [37]
union-find technique proposed by Rem [40]
CCL algorithm proposed in our paper using
scan strategy of ARUN and REMSP
Parallel
proposed in our paper
CCL algorithm suggested by [36]
CCL algorithm proposed in our paper using
scan strategy of CCLLRPC and REMSP
implementation
of AREMSP
union-find implementation thus allows us to process the pixels
of the image in any order. Therefore, we propose PAREMSP,
a parallel implementation of our proposed sequential two-pass
CCL algorithm AREMSP. For scalability, our algorithm in the
first pass, divides the image into equal proportions and executes
the scan strategy of ARUN algorithm along with REMSP
concurrently on each portion of the image. To merge the
provisional labels on the image boundary, we use the parallel
version of REMSP [38]. Our experiments show the scalability
of PAREMSP achieving speedups up to 20.1 using 24 cores
on shared memory architecture for an image of size 465.2 MB.
Additionally, the parallel algorithm does not make use of any
hardware specific routines, and thus is highly portable.
The remainder of this paper is organized as follows.
In section II, we provided related work on connected
component labeling. In section III, we propose our sequential
two-pass CCL algorithms CCLREMSP and AREMSP and
the parallel algorithm PAREMSP in section IV. We present
our experimental methodology and results in section V. We
conclude our work in section VI. The abbreviations used in
the paper and their brief description is given in Table I.
II. Related Work
As mentioned in [10], there exist different types of CCL
algorithms. Repeated pass or multi pass algorithm repeatedly
scans the image forward and backward alternatively to give
labels until no further changes can be made to the assigned
pixels [11], [12]. The algorithm in [10], which we call
as Suzuki's algorithm modifies the conventional multi pass
algorithm using one-dimensional table. There exists a parallel
implementation of Suzuki's algorithm using OpenMP in
[42]. According to experimental results in [42], the parallel
implementation gets maximum speedup of 2.5 on 4 threads.
In any two-pass algorithm, there are two steps in scanning
step: 1) examining neighbors of current pixel which already
assigned labels to determine label for the current pixel, 2)
storing label equivalence information to speed up the algorithm.
The algorithm in [36], which we refer to as CCLLRPC,
provides two strategies to improve the running time of the
algorithm. First strategy employs a decision tree, which
reduces the average number of neighbors accessed by a factor
of two. Second strategy replaces the conventional pointer
based union-find algorithm, which is used for storing label
equivalence, by adopting array based union-find algorithm that
2
uses less memory. The union-find algorithm is implemented
using Link by Rank and Path Compression technique.
The union-find data structure in [43] is replaced by a different
data structure to process label equivalence information. In this
algorithm, at any point, all provisional labels that are assigned
to a connected component found thus far during the first scan
are combined in a set S(r), where r is the smallest label and is
referred to as the representative label. The algorithm employs
rtable for storing representative label of a set, next to find
the next element in the set and tail to find the last element of
the set.
In another strategy, which we call ARUN, the first part of
scanning step employs a scanning technique, which processes
image two lines at a time and process two image pixels at a
time [37]. This algorithm uses the same data structure given in
[43] for processing label equivalence information. The scanning
technique reduces the number lines to be processed by half
thereby improving the speed of the two-pass CCL method.
In this paper, we provide two different implementations of
two-pass CCL algorithm. These two algorithms are different
in their first scan step. In the first implementation called
CCLREMSP, we have used the decision tree suggested by the
CCLLRPC algorithm for the first part of scanning step but
for the second part we have used REM's union-find approach
instead of Link by Rank and Path Compression technique.
The union-find algorithm maintains a collection of disjoint sets
where each set represents connected elements. [40] compares all
of the different variations of union-find algorithms over different
graph data sets and found that REM's implementation is best
among all the variations. Thus in our second implementation,
called AREMSP, we process the image lines two by two as
suggested by [37] but for the second step we use REMSP
instead of the data structure used by [37].
We have compared both of our proposed implementations
with CCLLRPC, RUN, and ARUN algorithms and find that
AREMSP performs best among all the algorithms. Finally we
have also provided a shared memory parallel implementation
of AREMSP called PAREMSP using OpenMP. We use the
parallel implementation of REMSP given in [38].
III. Proposed Algorithm
Throughout the paper, for an M × N image, we denote
image(a) to denote the pixel value of pixel a. We consider
binary images i.e. an image containing two types of pixels:
object pixel and background pixel. Generally, we consider
value of object pixel as 1 and value of background pixel as
0. The connected component labeling problem is to assign a
label to each object pixel so that connected object pixels have
the same label. In 2D images, there are two ways of defining
connectedness: 4-connectedness and 8-connectedness. In this
paper, we have only used 8-connectedness of a pixel.
A. CCLREMSP Algorithm
In CCLREMSP, we have used the decision tree suggested
in CCLLRPC (Figure 2) for scanning and REM's union-find
algorithm REMSP for storing label equivalence. The full
algorithm for CCLREMSP is given as Algorithm 1.
Algorithm 1 Pseudo-code for CCLREMSP
Input: 2D array image containing the pixel values
Output: 2D array label containing the final labels
1: function CCLREMSP(image)
2:
3:
4:
5:
6:
7:
8: end function
Scan CCLRemSP (image) (cid:46) Scan Phase of CCLREMSP
(cid:46) count is the max label assigned during Scan Phase
f latten(p, count)
for row in image do
(cid:46) Analysis Phase of CCLREMSP
(cid:46) Labeling Phase of CCLREMSP
for col in row do (cid:46) e is the current pixel to be labeled
label(e) ← p[label(e)]
In the scan step of CCLREMSP, we process image lines one
by one using the forward scan mask as shown in Figure 1a. We
have used the decision tree proposed by [36] for determining
the provisional label of current pixel e as we can reduce the
number of neighbors using decision tree. Instead of examining
all four neighbors of pixel, say e, i.e. a, b, c and d, we only
examine the neighbors according to a decision tree as shown
in Figure 2 [36]. Let label denote the 2D array storing the
labels and let p denote equivalence array then according to
CCLLRPC algorithm, three functions used by this decision
tree are defined as follows:
1). The one-argument copy function, copy(a), contains one
statement: label(e) = p(label(a))
2). The two-argument copy function, copy(c,a), contains one
statements: label(e) = merge(p, label(c), label(a))
3). The new label function sets count as label(e), appends
count to array p, and increments count by 1.
c
a
d
b
e
(a) Forward Scan
Mask for RemSP
c
b
a
e
d
g
f
(b)
Forward
Scan Mask for
ARemSP
Fig. 1: Forward Scan Mask
b
a
0
c
a
0
1
copy(a)
0
d
1
0
d
1
copy(b)
1
copy(c,a)
0
1
0
1
new label
copy(d)
copy(c)
copy(c,d)
Fig. 2: Decision tree suggested in CCLLRPC [36]
The implementation of Scan CCLRemSP is given as
3
Algorithm 4. However,
the implementation of MERGE
operation in our proposed algorithm REMSP is different from
that of in CCLLRPC. We have used the implementation of
union-find proposed by REM's for merge operation [40], [41].
REM's integrates the union operation with a compression
technique known as Splicing (SP ).
In the MERGE algorithm, if x and y are the nodes to be
merged then we set rootx to x and rooty to y. When rootx
is to be moved to p(rootx), firstly p(rootx) is stored in a
temporary variable z then p(rootx) is set to p(rooty), making
the subtree rooted at rootx a sibling of rooty and finally rootx
is set to z. The algorithm for MERGE is given as Algorithm
2. After the first step, we carry out the analysis phase using
the FLATTEN algorithm. In the FLATTEN algorithm, we give
smallest equivalent label of every connected component to
all the pixels belonging to that connected component. The
algorithm also generates consecutive labels. The algorithm for
FLATTEN is given as Algorithm 3.
if rootx = p[rootx]
Algorithm 2 Pseudo-code for merge [40]
Input: 1D array p and two nodes x and y
Output: The root of united tree
1: function MERGE(p,x,y)
rootx ← x, rooty ← y
2:
while p[rootx] (cid:54)= p[rooty] do
3:
if p[rootx] > p[rooty] then
4:
5:
p[rootx] ← p[rooty]
6:
return p[rootx]
7:
8:
9:
10:
11:
12:
13:
14:
15: end function
p[rooty] ← p[rootx]
return p[rootx]
return p[rootx]
then
z ← p[rooty], p[rooty] ← p[rootx], rooty ← z
else
z ← p[rootx], p[rootx] ← p[rooty], rootx ← z
if rooty = p[rooty]
then
Algorithm 3 Pseudo-code for flatten [36]
InOut: 1D array p containing the equivalence info
Input: Max value of provisional label count
1: function FLATTEN(p,count)
2:
3:
4:
5:
6:
7:
8:
9: end function
k ← 1
for i = 1 to count do
if p[i] < i then
p[i] ← p[p[i]]
p[i] ← k
k + +
else
B. AREMSP Algorithm
In AREMSP, we have used the decision tree suggested in
ARUN for scanning and REM's union-find algorithm for storing
label equivalence. The full algorithm for AREMSP is given as
Algorithm 5.
In the first scan step of AREMSP, we process an image two
lines at a time and two pixels at a time using the mask shown
in Figure 1b, which is suggested in [37].
else
copy(b)
if image(b) = 1 then
if image(c) = 1 then
if image(e) = 1 then
for row in image do
for col in row do
Algorithm 4 Pseudo-code for CCLREMSP Scan Phase
Input: 2D array image containing the pixel values
InOut: 2D array label containing the provisional labels and 1D array
p containing the equivalence info
Output: maximum value of provisional label in count
1: function SCAN CCLREMSP(image)
2:
3:
4:
5:
6:
7:
8:
9:
10:
11:
12:
13:
14:
15:
16:
17:
18:
19:
20:
21:
22:
23:
24:
25: end function
if image(a) = 1 then
if image(a) = 1 then
if image(d) = 1 then
if image(d) = 1 then
return count
new label
copy(c, a)
copy(c, d)
copy(a)
copy(d)
copy(c)
else
else
else
else
else
Algorithm 5 Pseudo-code for ARemSP
Input: 2D array image containing the pixel values
Output: 2D array label containing the final labels
1: function AREMSP(image)
2:
3:
4:
5:
6:
7:
8: end function
Scan ARemSP (image)
(cid:46) count is the max label assigned during Scan Phase
f latten(p, count)
for row in image do
(cid:46) Analysis Phase of RemSP
(cid:46) Labeling Phase of RemSP
for col in row do (cid:46) e is the current pixel to be labeled
label(e) ← p[label(e)]
(cid:46) Scan Phase of RemSP
We assign the label to both e and g simultaneously. If both
e and g are background pixels, then nothing needs to be done.
If e is a foreground pixel and there is no foreground pixel
in the mask, we assign a new provisional label to e and if
g is a foreground pixel, we will assign the label of e to g.
If there are foreground pixels in the mask, then we assign e
any label assigned to foreground pixels. In this case, if there
is only one connected component in the mask then there is
no need for label equivalence. Otherwise, if there are more
than one connected components in the mask and as they are
connected to e, all the labels of the connected components
are equivalent labels and hence need to be merged. For all
the cases, one can refer to [37]. However, our implementation
for the merge operation is different from [37]. We use the
implementation of union-find proposed by Rem [40], [41] for
the merge operation in AREMSP. Similar to CCLREMSP, we
use FLATTEN for analysis phase and generating consecutive
labels. The implementation of Scan ARemSP is given as
Algorithm 6.
4
else
else
merge(p, label(e), label(c))
if image(e) = 1 then
if image(d) = 0 then
merge(p, label(e), label(f ))
for row in image do
for col in row do
if image(a) = 1 then
label(e) ← label(a)
if image(c) = 1 then
if image(b) = 1 then
label(e) ← label(b)
if image(f ) = 1 then
if image(f ) = 1 then
label(e) ← label(f )
if image(a) = 1 then
merge(p, label(a))
if image(c) = 1 then
Algorithm 6 Pseudo-code for ARemSP Scan Phase
Input: 2D array image containing the pixel values
InOut: 2D array label containing the provisional labels and 1D array
p containing the equivalence info
Output: maximum value of provisional label in count
1: function SCAN AREMSP(image)
2:
3:
4:
5:
6:
7:
8:
9:
10:
11:
12:
13:
14:
15:
16:
17:
18:
19:
20:
21:
22:
23:
24:
25:
26:
27:
28:
29:
30:
31:
32:
33:
34:
35:
36:
37:
38:
39:
40:
41:
42:
43:
44:
45:
46:
47:
48: end function
if image(f ) = 1 then
label(g) ← label(f )
label(e) ← count,
p[count] ← count,
count + +
if image(c) = 1 then
label(e) ← label(c)
label(e) ← count,
p[count] ← count,
count + +
if image(d) = 1 then
label(g) ← label(d)
if image(g) = 1 then
label(g) ← label(e)
label(e) = label(d)
if image(b) = 0 then
if image(g) = 1 then
if image(c) = 1 then
merge(p, label(e), label(c))
merge(p, label(e), label(c))
return count
else
else
else
else
else
else
IV. Parallelizing AREMSP Algorithm
We now describe the parallel implementation of AREMSP
algorithm on a shared memory system. We make the assumption
about memory model as stated in OpenMP regarding the atomic
directive. We assume that memory read/write operations are
atomic and any operations issued concurrently by different
processors will be executed in some unknown sequential
order if no ordering constructs are being used. However, two
dependent operations issued by the same processor will always
be applied in the same order as they are issued.
numiter ← row/2 (cid:46) As we are processing 2 rows at a time
# pragma omp parallel
chunk ← numiter/numberof threads
size ← 2 × chunk
start ← start index of the thread
count ← start × col
# pragma omp for
Scan ARemSP (image)
# pragma omp for
for i = size to row − 1 do
if label(e) (cid:54)= 0 then
Algorithm 7 Pseudo-code for PARemSP
Input: 2D array image containing the pixel values
Output: 2D array label containing the final labels
1: function PAREMSP(image)
2:
3:
4:
5:
6:
7:
8:
9:
10:
11:
12:
13:
14:
15:
16:
17:
18:
19:
20:
21:
22:
23:
24:
25:
26: end function
merger(p, label(e), label(b))
if label(a) (cid:54)= 0 then
if label(c) (cid:54)= 0 then
i ← i + size
f latten(p, count)
for row in image do
for col in row do
label(e) ← p[label(e)]
if label(b) (cid:54)= 0 then
merger(p, label(e), label(a))
merger(p, label(e), label(c))
for col in row do
else
5
runs Scan Phase of AREMSP on it's chunk simultaneously.
We initialize the label to the start index of the thread for every
thread so that no two pixels in the image have the same label
after the first step. After the first step, each pixel is given a
provisional label. Next, the pixels at the boundary of each chunk
need to be merged to get the final labels. In the second step,
we merge the boundary pixels using parallel implementation of
Rem's Algorithm [38] which we call as MERGER. In MERGER,
if a thread wants to perform merging, it will first acquire the
necessary lock. Once it gets the lock, it will check whether
the node is still a root node. If yes, then the thread will set
the parent pointer and release the lock. On the other hand if
some other processor has altered the parent pointer so that the
node is no longer a root, the processor will release the lock
and continue executing the algorithm from its current position.
For complete reference, one can refer [38]. We implement
the parallel algorithm using OpenMP directives pragma omp
parallel and pragma omp for. The pseudo code of MERGER
is given as Algorithm 8. The pseudo code of PAREMSP is
given as Algorithm 7.
V. Experiments
For the experiments we used a computing node of Hopper, a
Cray XE6 distributed memory parallel computer. The node has
2 twelve-core AMD 'MagnyCours' 2.1-GHz processors and 32
GB DDR3 1333-MHz memory. Each core has its own L1 and
L2 caches, with 64 KB and 512 KB, respectively. One 6-MB L3
cache is shared between 6 cores on the MagnyCours processor.
All algorithms were implemented in C using OpenMP and
compiled with gcc.
Our test data set consists of four types of image data set:
Texture, Aerial, Miscellaneous and NLCD. First three data
sets are taken from the image database of the University of
Southern California.1 The fourth data set is taken from US
National Cover Database 2006.2 All of the images are converted
to binary images by MATLAB using im2bw(level) function
with level value as 0.5. The function converts the grayscale
image to a binary image by replacing all pixels in the input
image with luminance greater than 0.5 with the value 1 (white)
and replaces all other pixels with the value 0 (black). If the
input image is not a grayscale image, im2bw converts the input
image to grayscale, and then converts this grayscale image
to binary(Figure 3). However, note that our algorithm can be
easily extended to gray scale images.
Texture, Aerial and Miscellaneous data set contain images
of size 1 MB or less. NCLD data set contains images of size
bigger than 12 MB. The biggest image in the data set is 465.20
MB.
Firstly, we performed the experiment over all the sequential
algorithms. The experimental results are shown in Table II. In
the table, we have shown the minimum, maximum and average
execution time of all the four data sets. The execution time of
AREMSP is lowest among all the sequential algorithms. Thus
AREMSP is best among all the sequential algorithms.
Next, we show our results for the parallel algorithm
success ← 0
omp set lock(&(lock array[rootx]))
if rootx = p[rootx] then
p[rootx] ← p[rooty]
success ← 1
Algorithm 8 Pseudo-code for merger [38]
Input: 1D array p and two nodes x and y
Output: The root of united tree
1: function MERGER(p,x,y)
rootx ← x, rooty ← y
2:
while p[rootx] (cid:54)= p[rooty] do
3:
if p[rootx] > p[rooty] then
4:
if rootx = p[rootx] then
5:
6:
7:
8:
9:
10:
11:
12:
13:
14:
15:
16:
17:
18:
19:
20:
21:
22:
23:
24:
25:
26:
27: end function
return p[rootx]
break
break
else
z ← p[rootx], p[rootx] ← p[rooty], rootx ← z
if rooty = p[rooty] then
success ← 0
omp set lock(&(lock array[rooty]))
if root = p[rooty] then
p[rooty] ← p[rootx]
success ← 1
omp unset lock(&(lock array[rooty]))
if success = 1 then
omp unset lock(&(lock array[rootx]))
if success = 1 then
z ← p[rooty], p[rooty] ← p[rootx], rooty ← z
In PAREMSP, the image is divided row-wise into chunks of
equal size and given to the threads. In the first step, each thread
1http://sipi.usc.edu/database/
2http://dx.doi.org/10.1016/j.cageo.2013.05.014
(a) Original Image
(b) Binary Image
Fig. 3: Example of color image to binary image
Aerial
Miscellaneous
Texture
TABLE II: Comparison of various execution times[msec] for sequential algorithms
6
Image type
Aerial
Texture
Misc
NLCD
Min
Average
Average
Max
Min
Max
Min
Max
Min
Average
Average
Max
CCLLRPC
CCLRemSP
2.5
13.68
86.64
2.07
8.42
16.86
0.50
3.28
12.96
4.61
307.66
1307.27
2.48
13.25
80.90
2.06
8.20
16.18
0.49
3.21
12.81
4.46
299.55
1273.82
ARun
1.98
11.90
72.92
1.58
7.32
14.81
0.36
2.75
11.30
3.77
244.88
1036.52
ARemSP
1.95
11.86
70.17
1.53
7.27
14.47
0.36
2.74
11.20
3.75
242.59
1021.45
TABLE III: Images and their sizes [in MB]
Image name
image 1
image 2
image 3
image 4
image 5
image 6
Size
12
33
37.31
116.30
132.03
465.20
TABLE IV: Execution time [msec] of PAREMSP algorithm for various # threads
10
p
u
d
e
e
p
S
8
6
4
2
2
6
8
#Threads
16
24
Image type
Aerial
Texture
Fig. 4: Speedup for different images and different numbers of threads for Aerial,
Texture & Miscellaneous data set
Miscellaneous
NLCD
Min
Average
Max
Min
Max
Min
Max
Min
Average
Average
Average
Max
2
1.39
7.92
46.86
1.09
4.91
9.75
0.36
1.99
7.96
2.52
162.86
676.41
6
0.84
3.03
16.72
0.62
1.99
3.56
0.36
0.97
3.24
1.16
58.50
184.71
16
1.02
1.87
7.32
0.93
1.45
2.11
0.79
1.05
1.91
1.32
20.20
78.33
24
1.38
2.15
6.97
1.36
1.82
2.34
1.18
1.46
2.27
1.67
13.47
51.00
PAREMSP over all the images. We have shown the speedup
for data sets (except NCLD image data set) in Figure 4. The
minimum, maximum and average execution time of PAREMSP
for all the datasets is also shown in Table IV. We get a
maximum speedup of 10 in this case as the images are 1
MB or less in size. The speedup also decreases in some cases
as the number of threads increases. This case occurs when
the image size is small. As the number of threads increase,
each threads has less work, and therefore the thread creation
and termination overhead will affect the performance. Figure
5a-5b shows the speedup of the algorithm for NCLD image
data set. The size of the images are given in Table III. We
get a maximum speedup of 20.1 on 24 cores for image of
size 465.20 MB. Figure 5a shows the speedup for Phase-I of
PAREMSP i.e. the local computation and Figure 5b shows the
overall speedup (i.e. local + merge). We can see that there is
not significant difference between both speedups, implying that
merge operation does not have a significant overhead. Also
Figure 5 shows, speedup increases with image size.
Therefore, our parallel implementation is able to achieve
near linear speed for large data sets.
VI. Conclusion
In this paper, we presented two sequential CCL algorithms
CCLREMSP and AREMSP which are based on union-find
technique of REM's algorithm and scan strategies of ARUN
and CCLLRPC algorithms. CCLREMSP algorithm uses the
scan strategy of CCLLRPC algorithm whereas AREMSP
uses the scan strategy of ARUN algorithm. Based on the
experiments, we found out that AREMSP outperforms all the
other sequential algorithms. We also implement a portable
parallel
implementation of AREMSP for shared memory
computers with standard OpenMP directives. Our proposed
algorithm, PAREMSP, divides the image into equal proportions
and executes the scan. To merge the provisional labels on
the image boundary, we use the parallel version of REM's
algorithm. Our experimental results conducted on a shared
memory computer show scalable performance, achieving
speedups up to a factor of 20.1 when using 24 cores on data
set of size 465.20 MB. Thus, our parallel algorithm achieves
linear scaling for large fixed problem size while the number
of processing elements are increased.
Acknowledgment
This work is supported in part by the following grants:
NSF awards CCF-0833131, CNS-0830927,
IIS-0905205,
CCF-0938000, CCF-1029166, ACI-1144061, and IIS-1343639;
awards DE-FG02-08ER25848, DE-SC0001283,
DOE
p
u
d
e
e
p
S
25
20
15
10
5
0
0
5
10
15
# Threads
(a) local
7
p
u
d
e
e
p
S
25
20
15
10
5
0
20
25
0
5
10
15
# Threads
20
25
image 1
image 2
image 3
image 4
(b) local + merge
image 6
image 5
ideal
Fig. 5: Speedup for different images and different numbers of threads for NLCD data set
DE-SC0005309, DESC0005340, and DESC0007456; AFOSR
award FA9550-12-1-0458.
References
[1] Hussein M Alnuweiri and Viktor K Prasanna. Parallel architectures and
algorithms for image component labeling. IEEE Transactions on Pattern
Analysis and Machine Intelligence, 14(10):1014–1034, 1992.
[2] Rafael C Gonzales and RE Woods. Digital image processing, 1993.
[3] Pankaj K Agarwal, Lars Arge, and Ke Yi.
I/o-efficient batched
union-find and its applications to terrain analysis. In Proceedings of
the twenty-second annual symposium on Computational geometry, pages
167–176. ACM, 2006.
[4] Fu Chang, Chun-Jen Chen, and Chi-Jen Lu.
A linear-time
component-labeling algorithm using contour tracing technique. Computer
Vision and Image Understanding, 93(2):206–220, 2004.
[5] Hiroki Hayashi, Mineichi Kudo, Jun Toyama, and Masaru Shimbo. Fast
labelling of natural scenes using enhanced knowledge. Pattern Analysis
& Applications, 4(1):20–27, 2001.
[6] Qingmao Hu, Guoyu Qian, and Wieslaw L Nowinski.
Fast
connected-component labelling in three-dimensional binary images based
on iterative recursion. Computer Vision and Image Understanding,
99(3):414–434, 2005.
[7] Felipe Knop and Vernon Rego. Parallel labeling of three-dimensional
clusters on networks of workstations. Journal of Parallel and Distributed
Computing, 49(2):182–203, 1998.
[8] Alina N Moga and Moncef Gabbouj. Parallel image component labelling
with watershed transformation. Pattern Analysis and Machine Intelligence,
IEEE Transactions on, 19(5):441–450, 1997.
[9] Kuang-Bor Wang, Tsorng-Lin Chia, Zen Chen, and Der-Chyuan Lou.
Parallel execution of a connected component labeling operation on a
linear array architecture. J. Inf. Sci. Eng., 19(2):353–370, 2003.
[10] Kenji Suzuki,
Linear-time
connected-component labeling based on sequential local operations.
Computer Vision and Image Understanding, 89(1):1–23, 2003.
Isao Horiba, and Noboru Sugie.
[11] RM Haralick. Some neighborhood operators.
Computing, pages 11–35. Springer, 1981.
In Real-Time Parallel
[12] A Hashizume, R Suzuki, H Yokouchi, H Horiuchi, and S Yamamato. An
algorithm of automated rbc classification and its evaluation. Bio Medical
Engineering, 28(1):25–32, 1990.
[13] Toshiyuki Gotoh, Yoshiyuki Ohta, Masumi Yoshida, and Yoshiaki Shirai.
High-speed algorithm for component labeling. Systems and Computers
in Japan, 21(5):74–84, 1990.
[14] Toshiyuki Gotoh, Yoshiyuki Ohta, Masumi Yoshida, and Yoshio Shirai.
In Hague
Component labeling algorithm for video rate processing.
International Symposium, pages 217–224. International Society for Optics
and Photonics, 1987.
[15] Masatoshi Komeichi, Yoshiyuki Ohta, Toshiyuki Gotoh, Toshiya Mima,
In 1988 Intl
and Masumi Yoshida. Video-rate labeling processor.
Congress on Optical Science and Engineering, pages 69–76. International
Society for Optics and Photonics, 1989.
[16] Ronald Lumia. A new three-dimensional connected components
Computer Vision, Graphics, and Image Processing,
algorithm.
23(2):207–217, 1983.
[17] Ronald Lumia, Linda Shapiro, and Oscar Zuniga. A new connected
components algorithm for virtual memory computers. Computer Vision,
Graphics, and Image Processing, 22(2):287–300, 1983.
[18] Satoshi Naoi. High-speed labeling method using adaptive variable window
size for character shape feature. In IEEE Asian Conference on computer
vision, volume 1, pages 408–411, 1995.
[19] Azriel Rosenfeld. Connectivity in digital pictures. Journal of the ACM
(JACM), 17(1):146–160, 1970.
[20] Azriel Rosenfeld and John L Pfaltz. Sequential operations in digital
picture processing. Journal of the ACM (JACM), 13(4):471–494, 1966.
[21] Y Shirai. Labeling connected regions. Three-Dimensional Computer
Vision, pages 86–89, 1987.
[22] Michael B Dillencourt, Hannan Samet, and Markku Tamminen. A
general approach to connected-component labeling for arbitrary image
representations. Journal of the ACM (JACM), 39(2):253–280, 1992.
[23] Irene Gargantini and Zale Tabakman. Separation of connected component
In Proc. 12th Conf. Numerical
using linear quad-and oct-trees.
Mathematics and Computation, volume 37, pages 257–276, 1982.
[24] Jean Hecquard and Raj Acharya. Connected component labeling with
linear octree. Pattern recognition, 24(6):515–531, 1991.
[25] Hanan Samet. Connected component labeling using quadtrees. Journal
of the ACM (JACM), 28(3):487–501, 1981.
[26] Hanan Samet and Markku Tamminen. Computing geometric properties
of images represented by linear quadtrees. Pattern Analysis and Machine
Intelligence, IEEE Transactions on, (2):229–240, 1985.
[27] Hanan Samet and Markku Tamminen. Efficient component labeling of
images of arbitrary dimension represented by linear bintrees. Pattern
Analysis and Machine Intelligence, IEEE Transactions on, 10(4):579–586,
1988.
[28] Hanan Samet and Markku Tamminen. An improved approach to
connected component labeling of images. In International Conference
on Computer Vision And Pattern Recognition, pages 312–318, 1986.
[29] Markku Tamminen and Hanan Samet. Efficient octree conversion
In ACM SIGGRAPH Computer Graphics,
by connectivity labeling.
volume 18, pages 43–51. ACM, 1984.
[30] Prabir Bhattacharya. Connected component labeling for binary images
8
on a reconfigurable mesh architecture. Journal of Systems Architecture,
42(4):309–313, 1996.
on
[31] Alok Choudhary and Rajeev Thakur. Connected component labeling
on coarse grain parallel computers: an experimental study. Journal of
Parallel and Distributed Computing, 20(1):78–83, 1994.
[32] Daniel S. Hirschberg, Ashok K. Chandra, and Dilip V. Sarwate.
computers.
Computing
Communications of the ACM, 22(8):461–464, 1979.
components
connected
parallel
[33] M Manohar and HK Ramapriyan. Connected component labeling
of binary images on a mesh connected massively parallel processor.
Computer vision, graphics, and image processing, 45(2):133–149, 1989.
[34] David Nassimi and Sartaj Sahni. Finding connected components and
connected ones on a mesh-connected parallel computer. SIAM Journal
on computing, 9(4):744–757, 1980.
[35] Stephan Olariu, James L Schwing, and Jingyuan Zhang. Fast component
labelling and convex hull computation on reconfigurable meshes. Image
and vision computing, 11(7):447–455, 1993.
[36] Kesheng Wu, Ekow Otoo, and Kenji Suzuki. Optimizing two-pass
Pattern Analysis and
labeling algorithms.
connected-component
Applications, 12(2):117–135, 2009.
[37] Lifeng He, Yuyan Chao, and Kenji Suzuki. A new two-scan algorithm
for labeling connected components in binary images. In Proceedings of
the World Congress on Engineering, volume 2, 2012.
[38] Md Patwary, Mostofa Ali, Peder Refsnes, and Fredrik Manne. Multi-core
spanning forest algorithms using the disjoint-set data structure.
In
Parallel & Distributed Processing Symposium (IPDPS), 2012 IEEE
26th International, pages 827–835. IEEE, 2012.
[39] Chien-Wei Chen, Yi-Ta Wu, Shau-Yin Tseng, and Wen-Shan Wang.
Parallelization of connected-component labeling on tile64 many-core
platform. Journal of Signal Processing Systems, pages 1–15, 2013.
[40] Md Mostofa Ali Patwary, Jean Blair, and Fredrik Manne. Experiments on
union-find algorithms for the disjoint-set data structure. In Experimental
Algorithms, pages 411–423. Springer, 2010.
[41] Edsger Wybe Dijkstra, Edsger Wybe Dijkstra, Edsger Wybe Dijkstra,
and Edsger Wybe Dijkstra. A discipline of programming, volume 1.
prentice-hall Englewood Cliffs, 1976.
[42] Mehdi Niknam, Parimala Thulasiraman, and Sergio Camorlinga. A
parallel algorithm for connected component labelling of gray-scale
images on homogeneous multicore architectures. In Journal of Physics:
Conference Series, volume 256, page 012010. IOP Publishing, 2010.
[43] Lifeng He, Yuyan Chao, and Kenji Suzuki. A run-based two-scan labeling
Image Processing, IEEE Transactions on, 17(5):749–756,
algorithm.
2008.
|
1812.11774 | 1 | 1812 | 2018-12-31T12:23:27 | Tighter bounds for online bipartite matching | [
"cs.DS"
] | We study the online bipartite matching problem, introduced by Karp, Vazirani and Vazirani [1990]. For bipartite graphs with matchings of size $n$, it is known that the Ranking randomized algorithm matches at least $(1 - \frac{1}{e})n$ edges in expectation. It is also known that no online algorithm matches more than $(1 - \frac{1}{e})n + O(1)$ edges in expectation, when the input is chosen from a certain distribution that we refer to as $D_n$. This upper bound also applies to fractional matchings. We review the known proofs for this last statement. In passing we observe that the $O(1)$ additive term (in the upper bound for fractional matching) is $\frac{1}{2} - \frac{1}{2e} + O(\frac{1}{n})$, and that this term is tight: the online algorithm known as Balance indeed produces a fractional matching of this size. We provide a new proof that exactly characterizes the expected cardinality of the (integral) matching produced by Ranking when the input graph comes from the support of $D_n$. This expectation turns out to be $(1 - \frac{1}{e})n + 1 - \frac{2}{e} + O(\frac{1}{n!})$, and serves as an upper bound on the performance ratio of any online (integral) matching algorithm. | cs.DS | cs |
Tighter bounds for online bipartite matching
Uriel Feige ∗
January 1, 2019
Abstract
We study the online bipartite matching problem, introduced by Karp,
Vazirani and Vazirani [1990]. For bipartite graphs with matchings of size
n, it is known that the Ranking randomized algorithm matches at least
(1 −
e )n edges in expectation. It is also known that no online algorithm
e )n + O(1) edges in expectation, when the input
matches more than (1 −
is chosen from a certain distribution that we refer to as Dn. This upper
bound also applies to fractional matchings. We review the known proofs
for this last statement. In passing we observe that the O(1) additive term
(in the upper bound for fractional matching) is 1
n ), and that
this term is tight: the online algorithm known as Balance indeed produces
a fractional matching of this size. We provide a new proof that exactly
characterizes the expected cardinality of the (integral) matching produced
by Ranking when the input graph comes from the support of Dn. This
expectation turns out to be (1 −
n! ), and serves as an
upper bound on the performance ratio of any online (integral) matching
algorithm.
2e + O( 1
e )n + 1 −
2
e + O( 1
1
2 −
1
1
1
1 Introduction
Given a bipartite graph G(U, V ; E), where U and V are the sets of vertices and
E ∈ U × V is the set of edges, a matching M ⊂ E is a set of edges such that
every vertex is incident with at most one edge of M . Given a matching M , a
vertex is referred to as either matched or exposed, depending on whether it is
incident with an edge of M . A maximum matching in a graph is a matching
of maximum cardinality, and a maximal matching is a matching that is not a
proper subset of any other matching. Maximal matchings can easily be found
by greedy algorithms, and maximum matchings can also be found by various
polynomial time algorithms, using techniques such as alternating paths or linear
programming (see [9] and references therein). In every graph, the cardinality
of every maximal matching is at least half of that of the maximum matching,
because every matched edge can exclude at most two edges from the maximum
matching.
∗Department of Computer Science and Applied Mathematics, The Weizmann Institute.
[email protected]
1
For simplicity of notation, for every n we shall only consider the following
class of bipartite graphs, that we shall refer to as Gn. For every G(U, V ; E) ∈ Gn
it holds that U = V = n and that E contains a matching of size n (and
hence G has a perfect matching). The vertices of U will be denoted by ui (for
1 ≤ i ≤ n) and the vertices of V will be denoted by vi (for 1 ≤ i ≤ n). All
results that we will state for Gn hold without change for all bipartite graphs,
provided that n denotes the size of the maximum matching in the graph.
Karp, Vazirani and Vazirani [7] introduced an online version of the maximum
bipartite matching problem. This setting can be viewed as a game between two
players: a maximizing player who wishes the resulting matching to be as large
as possible, and a minimizing player who wishes the matching to be as small as
possible. First, the minimizing player chooses G(U, V ; E) in private (without the
maximizing player seeing E), subject to G ∈ Gn. Thereafter, the structure of G
is revealed to the maximizing player in n steps, where at step j (for 1 ≤ j ≤ n)
the set N (uj) ⊂ V of vertices adjacent to uj is revealed. At every step j, upon
seeing N (uj) (and based on all edges previously seen and all previous matching
decisions made), the maximizing player needs to irrevocably either match uj to
a currently exposed vertex in N (uj), or leave uj exposed.
There is much recent interest in the online bipartite matching problem and
variations and generalizations of it, as such models have applications for allo-
cation problems in certain economic settings, in which buyers (vertices of U )
arrive online and are interested in purchasing various items (vertices of V ). For
more details, see for example the survey by Metha [11].
An algorithm for the maximizing player in the online bipartite matching
setting will be called greedy if the only vertices of U that it leaves unmatched
are those vertices u ∈ U that upon their arrival did not have an exposed neighbor
(and hence could not be matched). It is not difficult to see that every non-greedy
algorithm A can be replaced by a greedy algorithm A′ that for every graph G
matches at least as many vertices as A does. Hence we shall assume that the
algorithm for the maximizing player is greedy, and this assumption is made
without loss of generality, as far as the results in this manuscript are concerned.
Every greedy algorithm (for the maximizing player) produces a maximal
matching, and hence matches at least half the vertices. For every deterministic
algorithm, the minimizing player can select a bipartite graph G (that admits
a perfect matching) that guarantees that the algorithm matches only half the
vertices. (Sketch: The first U
2 arriving vertices have all of V as their neighbors,
and the remaining U
2 vertices that the algorithm
matched with the first U
2 are neighbors only of the V
2 vertices.)
To improve the size of the matching beyond n
2 , Karp, Vazirani and Vazi-
rani [7] considered randomized algorithms for the maximizing player. Specifi-
cally, they proposed an algorithm called Ranking that works as follows. It first
selects uniformly at random a permutation π over the vertices V . Thereafter,
upon arrival of a vertex u, it is matched to its earliest (according to π) exposed
neighbor if there is one (and left unmatched otherwise). As the maximizing al-
gorithm is randomized (due to the random choice of π), the number of vertices
2
matched is a random variable, and we consider its expectation.
Let A be a randomized algorithm (such as Ranking) for the maximizing
player. As such, for every bipartite graph G it produces a distribution over
matchings. For a bipartite graph G ∈ Gn, we use the following notation:
• ρn(A, G) is the expected cardinality of matching produced by A when the
input graph is G.
• ρn(A, −) is the minimum over all G ∈ Gn of ρn(A, G). Namely, ρn(A, −) =
minG∈Gn[ρn(A, G)].
• ρn is the maximum over all A (randomized online matching algorithms
for the maximizing player) of ρn(A, −). Namely, ρn = maxA[ρn(A, −)].
(Showing that the maximum is attained is a technicality that we ignore
here.)
• ρ = inf n
ρn
n . Namely, ρ is the largest constant (independent of n) such that
ρn
ρ · n ≤ ρn for all n. (One might find a definition such as ρ = limn→∞
n
more natural, but it turns out that both definitions of ρ give the same
value, which will be seen to be 1 − 1
e .)
Karp, Vazirani and Vazirani [7] showed that ρn(Ranking, −) ≥ (1 − 1
e )n −
o(n), where e is the base of the natural logarithm (and (1 − 1
e ) ≃ 0.632). Un-
fortunately, that paper had only a conference version and not a journal version,
and the proof presented in the conference version appears to have gaps. Later
work (e.g., [12, 4, 2]), motivated by extensions of the online matching problem
to other problems such as the adwords problem, presented alternative proofs,
and also established that the o(n) term is not required. There have also been
expositions of simpler versions of these proofs. See [1, 10, 3], for example. Sum-
marizing this earlier work, we have:
Theorem 1 For every bipartite graph G ∈ Gn, the expected cardinality of the
matching produced by Ranking is at least (1 − 1
e )n. Hence ρn(Ranking, −) ≥
(1 − 1
e )n, and ρ ≥ 1 − 1
e ≃ 0.632.
Karp, Vazirani and Vazirani [7] also presented a distribution over Gn, and
showed that for every online algorithm, the expected size of the matching pro-
duced (expectation taken over random choice of graph from this distribution)
is at most (1 − 1/e)n + o(n). This distribution, that we shall refer to as Dn,
is defined as follows. Select uniformly at random a permutation τ over V . For
every j, the neighbors of vertex uj are {vτ (j), . . . , vτ (n)}. The unique perfect
matching M is the set of edges (uj, vτ (j)) for 1 ≤ j ≤ n.
To present the known results regarding Dn more accurately, let as extend
previous notation.
• ρn(A, Dn) is the expected cardinality of matching produced by A when
the input graph G is selected according to distribution Dn. (Hence expec-
tation is taken both over randomness of A and over selection from Dn.)
3
By definition, for every algorithm A, ρn(A, Dn) is an upper bound on
ρn(A, −).
• ρn(−, Dn) is the maximum over all A (randomized online algorithms for
the maximizing player) of ρn(A, Dn). Namely, ρn(−, Dn) = maxA[ρn(A, Dn)].
By definition, for every n, ρn(−, Dn) is an upper bound on ρn.
It is not hard to see (and was shown also in Lemma 13 of [7]) that for every
two greedy online algorithms A and A′ it holds that ρn(A, Dn) = ρn(A′, Dn).
As greedy algorithms are optimal among online algorithms, and Ranking is a
greedy algorithm, we have the following proposition.
Proposition 2 For Dn defined as above,
ρn(Ranking, Dn) = ρn(−, Dn) ≥ ρn
The result of [7] can be stated as showing that ρn(−, Dn) ≤ (1 − 1
e )n +
o(n). Later analysis (see for example [12], or the lecture notes of Kleinberg [8]
or Karlin [6]) replaced the o(n) term by O(1). Moreover, this upper bound
holds not only for online randomized integral algorithms (that match edges as a
whole), but also for online fractional algorithms (that match fractions of edges).
Let us provide more details.
A fractional matching for a bipartite graph G(U, V ; E) is a nonnegative
weight function w for the edges such that for every vertex u ∈ U we have
Pv∈N (u) w(u, v) ≤ 1, and likewise, for every vertex v ∈ V we havePu∈N (v) w(u, v) ≤
1. The size of a fractional matching is Pe∈E w(e). It is well known (see [9], for
example) that in bipartite graphs, the size of the maximum fractional matching
equals the cardinality of the maximum (integral) matching.
In the online bipartite fractional matching problem, as vertices of U arrive,
the maximizing player can add arbitrary positive weights to their incident edges,
provided that the result remains a fractional matching. We extend the ρ nota-
tion used for the integral case also to the fractional case, by adding a subscript
f . Hence for example, ρf,n(A, G) is the size of the fractional matching produced
by an online algorithm A when G ∈ Gn is the input graph.
It is not hard to see that in the fractional setting, randomization does not
help the maximizing player, in the sense that any randomized online algorithm
A for fractional matching can replaced by a deterministic algorithm A′ that
on every input graph produces a fractional matching of at least the same size.
(Upon arrival of vertex u, the fractional weight that A′ adds to edge (u, v) equals
the expected weight that A adds to this edge, where expectation is taken over
randomness of A.) Consequently, ρf,n ≥ ρn, and every upper bound on ρf,n is
also an upper bound on ρn.
The following theorem summarizes the known upper bounds on ρf,n, which
are also the strongest known upper bounds on ρn.
Theorem 3 For Dn as defined above, ρf,n(−, Dn) ≤ (1 − 1
quently, ρn(−, Dn) ≤ (1 − 1
e )n + O(1).
e )n + O(1). Conse-
4
The combination of Theorems 1 and 3 implies the following corollary:
Corollary 4 Using notation as above, ρ = 1 − 1
e . The Ranking
algorithm (which produces an integral matching) is asymptotically optimal (for
the maximizing player) for online bipartite matching both in the integral and
in the fractional case. The distribution Dn is asymptotically optimal for the
minimizing player, both in the integral and in the fractional case.
e and ρf = 1 − 1
In this manuscript, we shall be interested not only in the asymptotic ratios ρ
and ρf , but also in the exact ratios ρn and ρf,n. Every (integral) matching is also
a fractional matching, hence one may view Ranking also as an online algorithm
for fractional matching. As such, Ranking is easily seen not to be optimal
for some n. For example, when n = 4, tedious but straighforward analysis
shows that a different known algorithm referred to as Balance (see Section 2)
satisfies ρf,4(Balance, −) > ρf,4(Ranking, −) (details omitted). However, for
the integral case, it was conjectured in [7] that both Ranking and Dn are optimal
for every n. Namely, the conjecture is:
Conjecture 5 ρn = ρn(Ranking, Dn) for every n.
The above conjecture, though still open, adds motivation (beyond Proposi-
tion 2) to determine the exact value of ρn(Ranking, Dn). This is done in the
following theorem.
Theorem 6 Let the function a(n) be such that ρn(Ranking, Dn) = a(n)
for all
n!
n. Then a(n) = (n + 1)! − d(n + 1) − d(n), where d(n) is the number of derange-
ments (permutations with no fixed points) on the numbers [1, n]. Consequently,
ρn(−, Dn) = (1 − 1
e )n + 0.264, and this is also an
upper bound on ρn.
n! ) ≃ (1 − 1
e )n + 1 − 2
e + O( 1
The rest of this paper is organized as follows. In Section 2 we review a proof
of Theorem 3. In doing so, we determine the value of the O(1) term stated in
the theorem, and also show that the upper bound is tight. Hence we end up
proving the following theorem:
Theorem 7 For every n, Balance is the fractional online algorithm with best
approximation ratio, Dn is the distribution over graphs for which the approxi-
mation ratio is worst possible, and
ρf,n = ρf,n(Balance, Dn) = (1 −
1
e
)n +
1
2
−
1
2e
+ O(
1
n
) ≃ (1 −
1
e
)n + 0.316
In Section 3 we prove Theorem 6. The combination of Theorems 6 and 7 im-
plies that ρn < ρf,n for sufficiently large n. It also implies that ρf,n(Balance, Dn) >
ρf,n(Ranking, Dn) for sufficiently large n. Hence Proposition 2 does not extend
to online fractional matching.
In an appendix (Section A) we review a proof (due to [3]) of Theorem 1,
and derive from it an upper bound of (1 − 1
e on ρn(Ranking, Dn). This
last upper bound is weaker than the upper bounds of Theorems 6 and 7, but
its proof is different, and hence might turn out useful in attempts to resolve
Conjecture 5.
e )n + 1
5
1.1 Preliminaries -- MonotoneG
When analyzing ρn(Ranking, Dn) we shall use the following observation so as
to simplify notation. Because Ranking is oblivious to names of vertices, the
expected size of the matching produced by Ranking on every graph in the sup-
port of Dn is the same. Hence we shall consider one representative graph
from Dn, that we refer to as the monotone graph MonotoneG, in which γ
(in the definition of Dn) is the identity permutation. The monotone graph
G(U, V ; E) satisfies E = {(ui, vj ) j ≥ i}, and its unique perfect matching
is M = {(ui, vi) 1 ≤ i ≤ n}. Statements involving ρn(Ranking, Dn) will
be replaced by ρn(Ranking, M onotoneG), as both expressions have the same
value.
Likewise, the algorithm Balance is oblivious to names of vertices, and state-
ments involving ρf,n(Balance, Dn) will be replaced by ρf,n(Balance, M onotoneG).
2 Online fractional matchings
Let us present a specific online fractional matching algorithm that is often re-
ferred to as Balance, which is the natural fractional analog of an algorithm by
the same name introduced in [5]. Balance maintains a load ℓ(v) for every ver-
tex v ∈ V , equal to the sum of weights of edges incident with v. Hence at
all times, 0 ≤ ℓ(v) ≤ 1. Upon arrival of a vertex u with a set of neighbors
edges incident with u, maintaining the resulting loads as balanced as possi-
N (u), Balance distributes a weight of min[1, N (u) −Pv∈N (u) ℓ(v)] among the
ble. Namely, one computes a threshold t such that Pv∈N (u)ℓ(v)<t(t − ℓ(v)) =
min[1, N (u) − Pv∈N (u) ℓ(v)], and then adds fractional value t − ℓ(v) to each
edge (u, v) for those vertices v ∈ N (u) that have load below t.
We first present a proof of Theorem 3 based on previous work. The theorem
is restated below, with the additive O(1) term instantiated. Previous work
either did not specify the O(1) additive term (e.g., in [6]), or derived an O(1)
term that is not tight (e.g., in [8]).
Theorem 8 For every n it holds that
ρf,n(−, Dn) = (1 −
1
e
)n +
1
2
−
1
2e
+ O(
1
n
) ≃ (1 −
1
e
)n + 0.316
Moreover, ρf,n(−, Dn) = ρf,n(Balance, Dn).
Proof. For all graphs in the support of Dn, the size of the fractional matching
produced by Balance is the same (by symmetry). Hence for simplicity of no-
tation, consider the fractional matching produced by Balance when the input
graph is the monotone graph MonotoneG (see Section 1.1). It is not hard to see
that when vertex ui arrives, Balance raises the load of each vertex in {vi, . . . , vn}
by
n−i+1 ≤ 1. There-
after, when vertex uk+1 arrives, Balance can raise the load of its n − k neighbors
n−i+1 . This can go on until the largest k satisfying Pk
i=1
1
1
6
n−i+1 to 1. Hence altogether the size of the fractional matching is
i=1
n−i+1 ), for k as above.
1
i=1
from Pk
precisely k + (n − k)(1 −Pk
number Hn = Pn
1
The value of k can be determined as follows. It is known that the harmonic
n2 ), where γ ≃ 0.577 is the
Euler-Mascheroni constant. k is the largest integer such that Hn − Hn−k ≤ 1.
Defining α , n−k
i satisfies Hn = ln n+γ+ 1
2n +O( 1
i=1
n , we have that
1
Hn−Hn−k = ln n+γ+
1
2n
+O(
1
n2 )−ln αn−γ−
1
2αn
+O(
1
n2 ) = ln
1
α
−
1
α − 1
2n
+O(
1
n2 )
e (and temporarily ignoring the fact that in this case k = (1− 1
Choosing α = 1
is not an integer), we get that Hn − Hn−k = 1 − e−1
matching is then
e )n
n2 ). The size of a
2n + O( 1
(1 −
1
e
)n +
n
e
(
e − 1
2n
+ O(
1
n2 )) = (1 −
1
e
)n +
1
2
+
1
2e
+ O(
1
n
)
as desired.
The fact that k = (1 − 1
e )n above was not an integer requires that we round
k down to the nearest integer. The effect of this rounding is bounded by the
effect of changing the number of neighbors available to uk and to uk+1 by one
(compared to the computation without the rounding). Given that the number
of neighbors is roughly n
e , the overall effect on the size of the fractional matching
is at most O( 1
We conclude that ρf,n(Balance, Dn) = (1− 1
n ), implying that
2 + 1
ρf,n(−, Dn) ≥ (1 − 1
n ). In remains to show that ρf,n(−, Dn) ≤
(1 − 1
n ). This follows because Balance is the best possible
online algorithm (for fractional bipartite matching) against Dn. Let us provide
more details.
e )n + 1
2e + O( 1
2e + O( 1
2e +O( 1
e )n + 1
e )n+ 1
2 + 1
2 + 1
n ).
Given an input graph from the support of Dn, we shall say that a vertex
v ∈ V is active in round i if it is a neighbor of ui.
Initially all vertices are
active, and after every round, one more vertex (chosen at random among the
active vertices) becomes inactive, and remains inactive forever. Let a(i) denote
the number of active vertices at the beginning of round i, and note that a(i) =
n − i + 1. Consider an arbitrary online algorithm. Let L(i) denote the average
load of the active vertices at the beginning of round i. Then in round i, the
1
average load first increases by at most
a(i) (as long as it does not exceed 1) by
raising weights of edges, and thereafter, making one vertex inactive keeps the
average load unchanged in expectation (over choice of input from Dn). Hence
in expectation, in every round, the average load does not exceed the value of
the average load obtained by Balance. This means that in every round, in
expectation, the amount of unused load of the vertex that became inactive is
smallest when the online maximizing algorithm is Balance. Summing over all
rounds and using the linearity of expectation, Balance suffers the smallest sum
of unused load, meaning that it maximizes the final expected sum (over all V )
(cid:4)
of loads. The sum of loads equals the size of the fractional matching.
We now prove Theorem 7.
7
Proof.[Theorem 7] Given Theorem 8, it suffices to show that ρf,n(Balance, −) =
ρf,n(Balance, Dn), namely, that Dn is the worst possible distribution over input
graphs for the algorithm Balance. Moreover, given that Balance is oblivious to
the names of vertices, it suffices to show that MonotoneG is the worst possible
graph for Balance.
Let G(U, V ; E) ∈ Gn be a graph for which ρf,n(Balance, G) = ρf,n(Balance, −).
As Balance is oblivious to the names of vertices, we may assume that {(ui, vi)1 ≤
i ≤ n} is a perfect matching in G.
We use the notation N (w) to denote the set of neighbors of a vertex w in the
graph G. When running Balance on G, we use the notation m(i, j) to denote
the weight that the fractional matching places on edge (ui, vj) (and m(i, j) = 0
if (ui, vj) 6∈ E), and mi(j) to denote P1≤ℓ≤i m(uℓ, vj). Clearly, mi(j) is non-
decreasing in i. The size of the final fractional matching is m = Pn
j=1 mn(j).
When referring to a graph G′, we shall use the notation N ′ and m′ instead of
N and m.
An edge (ui, vj) with j < i is referred to as a backward edge.
Proposition 9 Without loss of generality, we may assume that G has no back-
ward edges. Hence mi(j) = mj(j) for all i > j.
Proof. Suppose otherwise, and let i be largest so that ui has backward edges.
Modify G by removing all backward edges incident with ui, thus obtaining a
graph G′. Compare the performance of Balance against the two graphs, G and
G′. On vertices u1, . . . , ui−1, both graphs produce the same fractional matching.
The extent to which ui is matched is at least as large in G as it is in G′ (because
also backward edges may participate in the fractional matching). Moreover, for
every vertex vj for i < j ≤ n, it holds that m′
i(j) ≥ mi(j). It follows that for
every vertex uℓ for ℓ > i, its marginal contribution to the fractional matching
in G is at least as large as its marginal contribution in G′. Hence the fractional
matching produced by Balance for G′ is not larger than that produced for G.
Repeating the above argument, all backward edges can be eliminated from G
(cid:4)
without increasing the size of the fractional matching.
Lemma 10 Without loss of generality we may assume that:
1. mi(i) ≤ mj(j) (or equivalently, mn(i) ≤ mn(j)) for all i < j.
2. mi(i) ≥ mi(j) for all i and j.
Proof. We first present some useful observations. For 1 ≤ i < n, consider
the set N (ui) of neighbours of ui in G (and recall that vi ∈ N (ui), and that
there are no backward edges). Then without loss of generality we may assume
that mi(i) ≥ mi(j) for all vj ∈ N (ui). This is because if there is some vertex
vj ∈ N (ui) with mi(j) > mi(i), then it must hold (by the properties of Balance)
that m(i, j) = 0. Hence the run of Balance would not change if the edge (ui, vj)
is removed from G (and then vj 6∈ N (ui)).
8
fractional matching in G′ (which is Pn
Moreover, we may assume that mi(i) = mi(j) for all vj ∈ N (ui). Suppose
otherwise. Then for vj ∈ N (ui) with smallest mi(j), modify G to a graph G′
as follows. For all ℓ < i, make uℓ a neighbor of vi iff it was a neighbor of vj,
and make uℓ a neighbor of vj iff it was a neighbor of vi. The final size of the
n(j)) cannot be larger than in G.
This is because m′
i(j) > mi(j) and for ℓ 6= j satisfying ℓ > i it
holds that m′
i(ℓ) = mi(ℓ). Moreover, as mi(i) < mi(j) ≤ 1, ui is fully matched
in G and hence also in G′, so the total size of fractional matching after step i
is the same in both graphs. Thereafter, the marginal increase of the fractional
matching at each step cannot be larger in G′ than it is in G.
i(i) < mi(i), m′
j=1 m′
By the same arguments as above we may assume that mi+1(i + 1) = mi+i(j)
for all vj ∈ N (ui+1).
Suppose now that item 1 fails to hold. Then for some 1 ≤ i ≤ n − 1 it holds
that mi(i) > mi+1(i + 1). Vertices ui and ui+1 cannot have a common neighbor
because if they do (say, vℓ) it holds that mi+1(i+1) = mi+1(ℓ) ≥ mi(ℓ) = mi(ii).
Hence we may exchange the order of ui and ui+1 (and likewise vi and vi+1)
without affecting the size of the fractional matching produced by Balance.
Repeating the above argument whenever needed we prove item 1 of the
lemma.
For j < i item 2 holds because mi(j) = mj(j) ≤ mi(i) (the last inequality
follows from item 1). For j > i item 2 holds because at the first point in time
ℓ ≤ i in which mℓ(j) = mi(j) it must be that mℓ(j) = mℓ(ℓ), and item 1 implies
(cid:4)
that mℓ(ℓ) ≤ mi(i).
It is useful to note that Lemma 10 implies that there is some round number
t such that for all ℓ ≥ t vertex vℓ is fully matched (namely, mn(ℓ) = 1), and for
every ℓ < t vertex vℓ is not fully matched (namely, mn(ℓ) < 1). As to vertices
in u, for ℓ < t vertex uℓ is fully matched, for ℓ > t vertex uℓ contributes nothing
to the fractional matching, and ut is either partly matched or fully matched.
Recalling that m denotes the size of the final fractional matching, we thus have
(for t as above):
m = t − 1 +Xj≥t
m(t, j)
(1)
At every step i, the contribution of vertex vi towards the fractional matching
is finalized at that step, namely, mn(i) = mi(i). Lemma 10 implies that for the
worst graph G, this vertex vi is the one with largest mi value at this given step.
Hence mi(i) = maxj≥i[mi(j)] and we have:
m =
n
Xi=1
mn(i) =
n
Xi=1
mi(i) =
n
Xi=1
max
j≥i
[mi(j)].
At this point it is intuitively clear why MonotoneG is the graph in Gn on
which Balance produces the smallest fractional matching. This is because with
MonotoneG, at each step i the fractional matching gets credited a value mi(i)
9
that is the average of the values mi(j) for j ≥ i, whereas for G its gets credited
the maximum of these values. Below we make this argument rigorous.
i(1) = 1
n Pn
j=1 mi(j). The remaining maxj≥1[m1(j)] − 1
Consider an alternative averaging process replacing algorithm Balance.
It
uses the same fractional matching as in Balance and the same m(i, j) values, but
maintains values m′
i(i) that may differ from mi(i). At round 1, instead of being
credited the maximum m1(1) = maxj≥1[m1(j)], the process is credited only the
average m′
j=1 m1(j)
is referred to as the slackness s(1). More generally, at every round i > 1,
instead of being credited by maxj≥i[mi(j)] at step i, the averaging process gets
credit from two sources. One part of the credit is the average
j=i mi(j),
where s(i) = maxj≥i[mi(j)] − 1
j=i mi(j) is the slackness generated at
round i. In addition, the process gets credit also for the slackness accumulated
in previous rounds ℓ < i, in such a way that each slackness variable s(ℓ) gets
distributed evenly among the n − ℓ rounds that follow it. Hence we set
n−i+1 Pn
n Pn
n−i+1 Pn
1
m′
i(i) =
1
n − i + 1
n
Xj=i
mi(j) +
i−1
Xℓ=1
s(ℓ)
n − ℓ
.
(2)
The averaging process continues until the first round t′ at which m′
which point m′
t′ (t′) ≥ 1, at
j(j) is set to 1 for all j ≥ t′, and the process ends. The size of the
i(i).
Computing m′ using the contributions of the vertices from U , for t′ as above,
we get that:
fractional matching associated with the averaging process is m′ = Pn
i=1 m′
m′ = t′ − 1 + Xj≥t′
m(t′, j)
(3)
Proposition 11 For the graph G, the size of the fractional matching produced
by the averaging process is no larger than that produced by Balance. Namely,
m′ ≤ m.
Proof. Compare Equations (1) and (3). If t′ = t then m′ = m, and if t′ < t
then m′ < m. Hence it suffices to show that the assumption t′ ≥ t implies that
t′ = t. This follows because mt(j) = 1 for all j ≤ t (as noted above), and so:
m′
t(t) =
1
n − t + 1
n
Xj=t
mt(j) +
t−1
Xℓ=1
s(ℓ)
n − ℓ
= 1 +
t−1
Xℓ=1
s(ℓ)
n − ℓ
≥ 1
where the last inequality holds because all slackness variables s(ℓ) are non-
(cid:4)
negative.
Proposition 12 For MonotoneG, running the averaging process and running
Balance are exactly the same process, giving m′(M onotoneG) = m(M onotoneG).
10
Proof. This is because when running Balance on MonotoneG, at every round
i we have that mi(i) = mi(j) for all j > i. Hence there is no difference between
(cid:4)
the average and the maximum of the mi(j) for j ≥ i.
Proposition 13 The size of the fractional matching produced by the averaging
process for graph G is not smaller than the size it produces for MonotoneG.
Namely, m′(G) ≥ m′(M onotoneG).
Proof. Running the averaging process on graph G, we claim that for every
round i < t′ we have that:
i(i) = Xk≤i
m′
1
n − k + 1
(4)
The equality can be proved by induction. For i = 1 both sides of the equality
n . For the inductive step, recalling Equation 2 one can infer that
are 1
m′
i+1(i + 1) =
1
n − i
((n − i + 1)m′
i(i) − m′
i(i) + 1)
where the +1 term is because i < t′. Likewise, the right hand side develops in
the same way:
1
1
1
n − k + 1
n − k + 1
1
=
n − k + 1
n − i
Xk≤i+1
(n − i + 1)Xk≤i
+ 1
The left hand side of Equation (4) concerns graph G. Observe that m′
i(i)
for MonotoneG exactly equals the right hand side of Equation (4). It follows
that the averaging process ends at the same step t′ both on the graph G and
on MonotoneG, and up to step t′ the accumulated fractional matching m′ is
identical. For rounds j ≥ t′ we have that m′
j(j) = 1 for G and it cannot be
(cid:4)
larger than 1 for MonotoneG, proving the proposition.
−Xk≤i
Combining the three propositions above we get that:
m(G) ≥ m′(G) ≥ m′(M onotoneG) = m(M onotoneG)
This completes the proof of Theorem 7.
(cid:4)
3 Online integral matching
The first part of Theorem 6 is restated in the following theorem (recall the
definition of the monotone graph MonotoneG in Section 1.1).
11
Theorem 14 Let the function a(n) be such that ρn(Ranking, M onotoneG) =
a(n)
for all n. Then a(n) = (n + 1)! − d(n + 1) − d(n), where d(n) is the number
n!
of derangements (permutations with no fixed points) on the numbers [1, n].
Proof. When the input is MonotoneG, then for every permutation π used by
Ranking, the matching M ′ produced satisfies the following two properties:
• All vertices in some prefix of U are matched, and then no vertices in the
resulting suffix are matched. This is because all neighbors of uj+1 are also
neighbors of uj, so if uj+1 is matched then so is uj.
• The order in which vertices of V are matched is consistent with the order
π (for those vertices that are matched -- some vertices of V may remain
unmatched). In other words, if two vertices vi and vj are matched and
π(i) < π(j), then the vertex u ∈ U matched with vi arrived earlier (has
smaller index) than the vertex u′ ∈ U matched with vj .
Some arguments in the proof that follows make use of the above properties,
without explicitly referring to them.
Fix n and MonotoneG as input. Let Πn denote the set of all permutations
over V . Hence Πn = n!. Ranking picks one permutation π ∈ Πn uniformly at
random. Recall our notation that π(i) is the rank of vi under π. We shall use
πi to denote the item of rank i in π (namely, πi = π−1(i)). For i ≤ n, let a(n, i)
denote the number of permutations π ∈ Πn under which πi is matched.
Proposition 15 For a(n) as defined in Theorem 14 and a(n, i) as defined
above, it holds that a(n) = Pn
i=1 a(n, i).
Proof. For a permutation π ∈ Πn, let x(π) denote the size of the greedy
matching produced when Ranking uses π and the input graph in MonotoneG.
Then by definition:
a(n) = Xπ∈Πn
x(π).
By changing the order of summation:
x(π) =
Xπ∈Πn
n
Xi=1
a(n, i).
Combining the above equalities proves the proposition.
(cid:4)
Proposition 15 motivates the study of the function a(n, i).
Lemma 16 The function a(n, i) satisfies the following:
1. a(n, 1) = n! for every n ≥ 1.
2. a(n, i) = a(n, i + 1) + a(n − 1, i) for every 1 ≤ i < n.
12
Proof. The first statement in the lemma holds because in every permutation π,
the item π1 is matched with u1. Hence it remains to prove the second statement.
Fixing n > 1 and i < n, consider the following bijection Bi : Πn −→ Πn,
where given a permutation π ∈ Πn, Bi(π) flips the order between locations i and
i + 1. Namely, Bi(π)i = πi+1 and Bi(π)i+1 = πi (we use Bi(π)i as shorthand
notation for (Bi(π))i). We compare the events that πi is matched by the greedy
matching when Ranking uses π with the event that Bi(π)i+1 is matched by the
greedy matching when Ranking uses Bi(π).
There are four possible events:
1. Both πi and Bi(π)i+1 are matched.
2. Neither πi nor Bi(π)i+1 are matched.
3. πi is matched but Bi(π)i+1 is not matched.
4. πi is not matched but Bi(π)i+1 is matched.
Though any of the first three events may happen, the fourth event cannot
possibly happen. This is because the item in location i + 1 in Bi(π) is moved
forward to location i in π, so if the greedy algorithm matches it (say to uj) in
Bi(π), then the greedy algorithm must match it (either to the same uj or to the
earlier uj−1) in π.
It follows that a(n, i) − a(n, i + 1) exactly equals the number of permutations
in which the third event happens. Hence we characterize the conditions under
which the third event happens. Let uj be the vertex matched with πi in π.
Up to the arrival of uj, the behavior of Ranking on Bi(π) and π is identical.
Thereafter, for uj not to be matched to Bi(π)i+1 = πi, it must be matched to
the earlier Bi(π)i. Thereafter, for uj+1 not to be matched to Bi(π)i+1, it must
be that Bi(π)i+1 is not a neighbor of uj+1. But Bi(π)i+1 = πi is a neighbor
of uj (it was matched to uj under π), and hence it must be that πi = vj.
Summarizing, the third event happens if and only if the permutation Bi(π)
comes from the following class Π, where permutations π ∈ Π are those that
have the property that πi is matched, and πi+1 = vj, for the same j for which
uj is the vertex matched with πi. Consequently, a(n, i) = a(n, i + 1) + Π.
To complete the proof of the lemma, it remains to show that Π = a(n−1, i).
Let Π′ ⊂ Πn−1 be the set of these permutations π′ ∈ Πn−1 under which
Ranking (when U = V = n − 1) matches the item π′
i.
Claim 17 For Π and Π′ as defined above it holds that Π = Π′.
Proof. We first show a mapping from Π to Π′. Given π ∈ Π, let vj = πi+1. To
obtain permutation π′ ∈ Πn−1 from π, remove vj from π, identify location k in
π with location k − 1 in π′ (for i + 2 ≤ k ≤ n), and identify item vℓ of π with
item vℓ−1 of π′ (for j + 1 ≤ ℓ ≤ n). We show now that π′ ∈ Π′ (namely, π′
i is
matched, when the input graph is MonotoneG with U = V = n − 1).
The vertices u1, . . . , uj−1 are matched to exactly the same locations in π′
and in π, because the only vertices whose indices were decremented had index
13
ℓ ≥ j +1, and are neighbors of u1, . . . , uj−1 both before and after the decrement.
Let vk = πi and note that k > j, because vk is matched to uj and it is not
vj = πi+1. Hence π′
i = vk−1 and it too is a neighbor of uj, because j ≤ k − 1.
Hence π′
i will be matched to uj.
Conversely, we have the following mapping from Π′ to Π. Given π′ ∈ Π′, let
i. To obtain permutation π ∈ Π from π′, identify
uj be the vertex matched π′
location k in π with location k − 1 in π′ (for i + 2 ≤ k ≤ n), identify item vℓ of
π with item vℓ−1 of π′ (for j + 1 ≤ ℓ ≤ n), and set πi+1 = vj. We show now
that π ∈ Π.
As in the first mapping, the vertices u1, . . . , uj−1 are matched to exactly the
same locations in π′ and in π. Let vk = π′
i and note that k ≥ j, because vk was
matched to uj. Hence πi = vk+1 is neighbor of uj, and will be matched to uj.
On the other hand, πi+1 = vj will not be matched because it is not a neighbor
of any of [uj+1, un]. Hence π ∈ Π.
Given the two mappings described above (one is the inverse of the other) we
(cid:4)
have a bijection between Π′ and Π, proving the claim.
The claim above implies that Π = Π′ = a(n − 1, i), and consequently that
(cid:4)
a(n, i) = a(n, i + 1) + a(n − 1, i), proving the lemma.
In passing, we note the following corollary.
Corollary 18 For a(n, i) and a(n) as defined above, a(n) = (n + 1)! − a(n +
1, n + 1).
Proof. Using item 1 of Lemma 16 we have that a(n + 1, 1) = (n + 1)!. Applying
item 2 of Lemma 16 iteratively for all 1 ≤ i ≤ n we have that a(n + 1, 1) −
i=1 a(n, i) = a(n).
Combining these three equalities we obtain a(n) = (n + 1)! − a(n + 1, n + 1), as
(cid:4)
desired.
i=1 a(n, i). Proposition 15 shows that Pn
a(n + 1, n + 1) = Pn
Corollary 18 can also be proved directly, without reference to Lemma 16.
See Appendix B for details.
To obtain expressions for the values a(n, i), let us introduce additional no-
tation. A fixpoint (or fixed point) in a permutation π is an item that does not
change its location under π (namely, π(i) = i). For n ≥ 1 and 1 ≤ i ≤ n define
d(n, i) be the number of permutations over [n] in which the only fixpoints (if any)
are among the first i items. For example, d(3, 1) = 3 due to the permutations
132 (only 1 is a fixed point) 231 (no fixpoints) and 312 (no fixpoints).
Lemma 19 The function d(n, i) satisfies the following:
1. d(n, n) = n! for every n ≥ 1.
2. d(n, i + 1) = d(n, i) + d(n − 1, i) for every 1 ≤ i < n.
14
Proof. d(n, n) denotes the number of permutations on [n] with no restrictions,
and hence d(n, n) = n!, which is the first statement of the lemma.
Consider now the second statement of the lemma. Let Πn,i denote the
set of permutations in which the only fixpoints (if any) are among the first
i items. Then the second statement asserts that Πn,i+1 = Πn,i + Πn−1,i.
The set Πn,i+1 can be partitioned in two. In one part i + 1 is not a fixpoint.
This part is precisely Πn,i. In the second part, i + 1 is a fixpoint. To specify
a permutation in this part we need to specify the location of the remaining
n − 1 items, where the only fixpoints allowed are among the first i items. The
number of permutations satisfying these constraints is Πn−1,i, by definition.
(cid:4)
Hence indeed Πn,i+1 = Πn,i + Πn−1,i, proving the lemma.
Corollary 20 For every n ≥ 1 and 1 ≤ i ≤ n it holds that a(n, i) = d(n, n +
1 − i).
Proof. The proof is by induction on n, and for every value of n, by induction
on i.
For the base case n = 1, necessarily i = 1 (and hence also n + 1 − i = 1)
and indeed we have a(1, 1) = 1 = d(1, 1). Fixing n > 1, the base case for i is
i = 1 (and n + 1 − 1 = n) and indeed we have that a(n, 1) = n! = d(n, n). For
the inductive step, consider a(n, i) with n > 1 and 1 < i ≤ n, and assume the
inductive hypothesis for n′ < n and the inductive hypothesis for n and i′ < i.
Then we have:
a(n, i) = a(n, i−1)−a(n−1, i−1) = d(n, n−i+2)−d(n−1, n−i+1) = d(n, n−i+1)
The first equality is by Lemma 16, the second equality is by the inductive
(cid:4)
hypothesis, and the third equality is by Lemma 19.
We can now complete the proof of Theorem 14. By Corollary 18 we have that
a(n) = (n + 1)! − a(n + 1, n + 1). By Corollary 20 we have that a(n + 1, n + 1) =
d(n + 1, 1). By definition, d(n + 1, 1) is the number of permutations on [n + 1]
in which only item 1 is allowed to be a fixpoint. This number is precisely
d(n + 1) + d(n) (where d(j) are the derangement numbers), where the term
d(n + 1) counts those permutations in which there is no fixpoint, and the term
(cid:4)
d(n) counts those permutations in which item 1 is the only fixpoint.
The second part of Theorem 6 is restated in the following Corollary.
Corollary 21 For every n,
ρn(Ranking, M onotoneG) = (1 +
1
e
)n + (1 −
2
e
) + ν(n)
where ν(n) < 1
n! .
15
Proof. Theorem 14 shows that a(n) = (n + 1)! − d(n + 1) − d(n), where d(n) are
the derangement numbers. It is known that d(n) = n!
e rounded to the nearest
integer. Hence d(n) − n!
e < 1. Hence
a(n) − (1 − 1
e )(n + 1)
by (1 − 1
(cid:4)
e < 1. Dividing by n! and replacing (1 − 1
e )(n + 1)! − n!
e the corollary is proved.
e < 1
2 and d(n + 1) + d(n) − (n+1)!
e − n!
e )n + 1 − 1
3.1 Some related sequences
To illustrate the values of some of the parameters involved in the proof of The-
orem 14, consider a triangular table T where row n has n columns. The en-
tries (for 1 ≤ i ≤ n) are d(n, i), as defined prior to Lemma 19. Recall that
d(n, i) = a(n, n + 1 − i), hence the table also provides the a(n, i) values. We ini-
tialize the diagonal of the table by d(n, n) = n!. Thereafter we fill the remaining
cells of table row by row, by using the relation d(n, i) = d(n, i + 1) − d(n − 1, i),
i=1 d(n, i)
by summing up each row. The table below shows the computation of a(n) for
n ≤ 6.
implied by Lemma 19. Finally, compute a(n) = Pn
i=1 a(n, i) = Pn
n
1
2
3
4
5
6
d(n,1)=a(n,n)
d(n,2)
d(n,3)
d(n,4)
d(n,5)
d(n,6)
1
1
3
11
53
309
2
4
14
64
362
6
18
78
426
24
96
504
120
600
720
a(n)
1
3
13
67
411
2921
The table T is identical in its definition to Sequence A116853, named Dif-
ference triangle of factorial numbers read by upward diagonals, in The Online
Encyclopedia of Integer Sequences [13]. The row sums (and hence a(n)) in this
table give Sequence A180191 (with an offset of 1 in the value of n), named
Number of permutations of [n] having at least one succession. The first col-
umn (which equals a(n, n)) is the sequence A000255. These relations between
a(n) and the various sequences in [13] helped guide the statement and proof of
Theorem 14.
The derangement numbers d(n) (which form the sequence A000166) can be
easily computed by the recurrence d(n) = n · d(n − 1) + (−1)n (due to Euler).
The table below shows the computation of a(n) = (n + 1)! − d(n + 1) − d(n) for
n ≤ 7.
16
n
1
2
3
4
5
6
7
8
n!
1
2
6
24
120
720
5040
40320
d(n)
0
1
2
9
44
265
1854
14833
a(n)
1
3
13
67
411
2921
23633
Acknowledgements
The work of the author is supported in part by the Israel Science Foundation
(grant No. 1388/16). The results reported in this manuscript were obtained
in preparation for a talk given at the event Building Bridges II: Conference to
celebrate the 70th birthday of Laszlo Lovasz, Budapest, July 2018. I thank several
people whose input helped shape this work. Alon Eden and Michal Feldman
directed me to the proof presented in [3], which is the one presented here (in
the appendix) for Theorem 1. Thomas Kesselheim and Aranyak Mehta directed
me to additional relevant references. The statement and proof of Theorem 14
were based on noting some numerical coincidences between the values of a(n)
for small n and sequences in The Online Encyclopedia of Integer Sequences [13].
Dror Feige wrote a computer program that computes a(n), which made these
numerical coincidences evident. Alois Heinz offered useful advice as to how to
figure out proofs for various identities claimed in [13].
References
[1] Benjamin E. Birnbaum, Claire Mathieu: On-line bipartite matching made
simple. SIGACT News 39(1): 80 -- 87 (2008).
[2] Nikhil R. Devanur, Kamal Jain, Robert D. Kleinberg: Randomized Primal-
Dual analysis of RANKING for Online BiPartite Matching. SODA 2013:
101 -- 107.
[3] Alon Eden, Michal Feldman, Amos Fiat, Kineret Segal:
An
Economic-Based Analysis of RANKING for Online Bipartite Matching.
https://arxiv.org/abs/1804.06637.
[4] Gagan Goel, Aranyak Mehta: Online budgeted matching in random input
models with applications to Adwords. SODA 2008: 982 -- 991.
[5] Bala Kalyanasundaram, Kirk Pruhs: An optimal deterministic algorithm
for online b-matching. Theor. Comput. Sci. 233(1-2): 319 -- 325 (2000).
17
[6] Anna Karlin. Online
bipartite matching,
on Randomized Algorithms
course
sis,
https://courses.cs.washington.edu/courses/cse525/13sp/scribe/lec6.pdf
scribes Alex Polozav
and Daryl Hansen,
Spring
and
notes
lecture
in
Probabilistic Analy-
2013.
[7] Richard M. Karp, Umesh V. Vazirani, Vijay V. Vazirani: An Optimal
Algorithm for On-line Bipartite Matching. STOC 1990: 352 -- 358.
[8] Robert D. Kleinberg. Online
bipartite matching
lecture note
http://www.cs.cornell.edu/courses/cs6820/2012fa/
on Analysis
in course
of Algorithms, Fall
algorithms,
2012.
[9] L. Lovasz, M. D. Plummer. Matching Theory. Elsevier 1986.
[10] Claire Mathieu.
Professor
Blog:
CS's
A
the Ranking
analysis
dual
http://teachingintrotocs.blogspot.co.il/2011/06/primal-dual-analysis-of-ranking.html
algorithm.
June
of
A
25,
primal-
2011.
[11] Aranyak Mehta. Online matching and ad allocation. Foundations and
Trends in Theoretical Computer Science, 8(4):265 -- 368, 2013.
[12] Aranyak Mehta, Amin Saberi, Umesh V. Vazirani, Vijay V. Vazirani: Ad-
Words and generalized online matching. J. ACM 54(5): 22 (2007).
[13] N. J. A. Sloane, editor, The On-Line Encyclopedia of Integer Sequences,
published electronically at https://oeis.org.
A A performance guarantee for Ranking
For completeness, we review here a proof of Theorem 1. The proof that we
present uses essentially the same mathematical expressions as the proof pre-
sented in [2]. A simple presentation of the proof of [2] appeared in a blog post
of Claire Mathieu [10] (with further slight simplifications made possible by a
comment provided there by Pushkar Tripathi). We shall give an arguably even
simpler presentation, due to Eden, Feldman, Fiat and Segal [3]. The proofs
in [2, 10] make use of linear programming duality. The proof below is based on
an economic interpretation, and a proof technique that splits welfare into the
sum of utility and revenue. These last two terms turn out to be scaled versions
of the dual variables used in [2, 10], but the proof does not need to make use of
LP duality.
Proof.[Theorem 1] Fix an arbitrary perfect matching M in G. Given a vertex
v ∈ V , we use M (v) to denote the vertex in U matched with v under M .
Recall that Ranking chooses a random permutation π over V . Equivalently,
we may assume that every vertex vi ∈ V chooses independently uniformly at
random a real valued weight wi ∈ [0, 1], and then the vertices of V are sorted in
order of increasing weight (lowest weight first). This gives a random permutation
π. The same permutation π is also obtained if each weight wi is replaced
by a "price" pi = ewi−1 and vertices are sorted by prices (because ex−1 is a
18
monotonically increasing function in x). Observe that pi ∈ [ 1
e , 1], though it
is not uniformly distributed in that range. The expected price that Ranking
assigns to an item is:
E[pi] = Z 1
0
ewi−1dwi =
1
e
(e − 1) = 1 −
1
e
(5)
It is convenient to think of the vertices of U as buyers and the vertices of
V as items. Suppose that given G(U, V ; E), each vertex (buyer) u ∈ U desires
only items v ∈ V that are neighbors of u (namely, u desires v iff (u, v) ∈ E), is
willing to pay 1 for any such item, and wishes to buy exactly one item. The seller
holding the items is offering to sell each item vi for a price of pi. Then given G,
the matching produced by executing the Ranking algorithm is the same as the
one that would be produced in a setting in which each buyer uj, upon arrival,
buys its cheapest exposed desired item, if there is any. If pi is the price of the
purchased item vi, then the revenue that the seller extracts from the sale of vi
to uj is r(vi) = pi, whereas the utility that the buyer extracts is y(ui) = 1 − pi.
Consequently, the revenue plus utility extracted from a sale is 1, and the total
revenue extracted from all sales plus the total utility sum up to exactly the
cardinality of the matching.
To lower bound the expected cardinality of the matching, we consider each
edge (M (vi), vi) ∈ M separately, and consider the expectation E[r(vi)+y(M (vi))],
where expectation is taken over the choice of π. Using the linearity of the ex-
pectation, we will have that ρn(Ranking, G) = Pvi∈V E[r(vi) + y(M (vi))].
Lemma 22 For every vi ∈ V it holds that E[r(vi) + y(M (vi))] ≥ 1 − 1
e . More-
over, this holds even if expectation is taken only over the choice of random
weight wi (and hence of random price pi) of item vi, without need to consider
other aspects of the random permutation π.
Proof. Fix an arbitrary graph G(U, V ; E) ∈ Gn, an arbitrary perfect matching
M , and arbitrary prices pj ∈ [ 1
e , 1] for all items vj 6= vi. The price pi for item vi is
set at random. Let M ′ denote the greedy matching produced by this realization
of the Ranking algorithm (where each buyer upon its arrival is matched to the
exposed vertex of lowest price among its neighbors, if there is any). Suppose as
a thought experiment that item vi is removed from V , and consider the greedy
matching M ′
−i that would have been produced in this setting. Let p denote the
price of the item in V matched to M (vi) under M ′
−i, and set p = 1 if M (vi) is
left unmatched under M ′
−i. Now we make two easy claims.
1. If pi < p, then vi is matched in M ′. This follows because at the time
that M (vi) arrived, either vi was already matched (as desired), or it was
available for matching with M (vi) and preferable (in terms of price) over
all other items that M (vi) desires (as all have price at least p > pi).
2. The utility of M (vi) in M ′ satisfies y(M (vi)) ≥ 1−p. This follows because
−i the utility of M (vi) is 1 − p, and under the greedy algorithm
under M ′
19
considered, the introduction of an additional item (the item vi when con-
sidering M ′) cannot decrease the utility of any agent. (At every step of
the arrival process, the set of exposed vertices under M ′ contains the set
of exposed vertices under M ′
−i, and one more vertex.)
Using the above two claims and taking z to be the value satisfying p = ez−1,
we have:
E[y(M (vi))+r(vi)] ≥ 1−p+P r[pi < p]pi = 1−ez−1+Z z
wi=0
ewi−1dwi = 1−
ez
e
+
ez − 1
e
= 1−
1
e
This completes the proof of Lemma 22.
Using the linearity of the expectation, we have that
ρn(Ranking, G) = Xvi∈V
E[r(vi) + y(M (vi))] ≥ (1 −
1
e
)n
This completes the proof of Theorem 1.
(cid:4)
(cid:4)
One can adapt the proof presented above to the special case in which the
input graph is MonotoneG (or more generally, comes from the distribution Dn).
In this case one can upper bound the slackness involved in the proof of Theo-
rem 1, and infer the following theorem.
Theorem 23 For every n it holds that ρn(Ranking, M onotoneG) ≤ (1 − 1
1
e .
e )n+
Proof. Recall the two properties mentioned in the beginning of the proof of
Theorem 14. Recall also that the analysis of Ranking in the proof of Theorem 1
(within Lemma 22) involved the matching M ′ and other matchings M ′
−i, and
two claims. Let us analyse the slackness involved in these claims when the
input is the monotone graph. The claims are restated with M (vi) replaced by
ui, because for the monotone graph M (vi) = ui.
The first claim stated that if pi < p, then vi is matched in M ′. When the
input is the monotone graph, then a converse also holds: if pi > p, then vi is
not matched in M ′. This follows because up to the time that ui arrives and is
matched, only vertices of V priced at most p are matched, and thereafter, no
other vertex in U desires vi. The event that pi = p has probability 0. Hence
there is no slackness involved in the first claim -- it is an if and only if statement.
The second claim stated that the utility of ui in M ′ is y(ui) ≥ 1 − p. This
inequality is not tight. Rather, the utility of ui in M ′
−i is 1 − p, and y(ui)
is not smaller. Let us quantify the slackness involved in this inequality by
introducing slackness variables s(u). For a vertex u ∈ U we shall use the
notation y(u) to denote the utility of u under Ranking, and y−v(u) for the
utility of u when vertex v ∈ V is removed. The slackness s(ui) of vertex ui is
defined as s(ui) = y(ui) − y−vi(ui).
20
Lemma 24 For the monotone graph and an arbitrary vertex uj ∈ U , the ex-
pected utility of uj (expectation taken over choices of wi for all 1 ≤ i ≤ n by
the Ranking algorithm) is identical in the following two settings: when vj is
removed, and when vn is removed. Namely, E[y−vj (uj)] = E[y−vn (uj)].
Proof. Both vj and vn are neighbors of all vertices uk arriving up to uj (for
1 ≤ k ≤ j). Hence whichever of the two vertices, vj or vn, is removed, the dis-
tributions of the outcomes of Ranking on the first j arriving vertices (including
(cid:4)
uj) are the same.
As a consequence of Lemma 24 we deduce that for the monotone graph,
the expected slackness of every vertex u ∈ U satisfies E[s(u)] = E[y(u)] −
E[y−vn (u)].
Lemma 25 For the monotone graph and arbitrary setting of prices for the
items (as chosen at random by Ranking), Pu∈U s(u) ≤ 1 − pn. Consequently,
Pu∈U E[s(u)] ≤ 1
e , where expectation is taken over choice of weights wi for
vertices in V .
Proof. Fix the prices pi (hence π). Let u1, . . . , uk be the vertices of U matched
under Ranking, and let m(u1), . . . , m(uk) be the vertices in V to which they
are matched. Observe that the prices p(m(ui)) (where 1 ≤ i ≤ k) of these
vertices form a monotonically increasing sequence. Necessarily, vn is one of the
matched vertices, because it is a neighbor of all vertices in U . Let j be such
that vn = m(uj).
Consider now what happens when vn is removed. The vertices u1, . . . , uj−1
are matched to m(u1), . . . , m(uj − 1) as before. As to the vertices uj, . . . , uk−1,
they can be matched to m(uj+1), . . . , m(uk), hence the algorithm will match
them to vertices of no higher price. Specifically, for every i in the range j ≤
i ≤ k − 1, vertex ui will be matched either to m(ui+1) or to an earlier vertex,
though not earlier than m(ui). The vertex uk may either be matched or be
left unmatched. For simplicity of notation, we say that uk is matched to either
m(uk+1) or to an earlier vertex, where m(uk+1) is an auxiliary vertex of price 1
than indicates that uk is left unmatched.
Note that:
and that:
y(u) =
Xu∈U
k
Xi=1
y(ui) = k −
k
Xi=1
p(m(ui))
y−vn(u) =
Xu∈U
k
Xi=1
y−vn(ui) ≥ k −
j−1
Xi=1
p(m(ui)) −
k+1
Xi=j+1
p(m(ui))
Hence we have that:
21
Xu∈U
s(u) = Xu∈U
y(u) − Xu∈U
y−vn(u) ≤ p(m(uk+1)) − p(vn)
Finally, noting that p(m(uk+1)) ≤ 1 and that E[p(vn)] = 1 − 1
e (see Equa-
(cid:4)
tion (5)), the lemma is proved.
As in the proof of Theorem 1 we have:
E[y(ui) + r(vi)] = 1 − p + s(ui) + P r[pi < p]pi = 1 −
1
e
+ s(ui)
Using the linearity of the expectation and Lemma 25 we have that:
ρn(Ranking, M onotoneG) = Xvi∈V
E[r(vi)+y(ui)] = (1−
1
e
)n+Xu∈U
E[s(u)] ≤ (1−
1
e
)n+
1
e
This completes the proof of Theorem 23.
(cid:4)
B An alternative proof of a combinatorial iden-
tity
We present a proof of Corollary 18 that does not make use of Lemma 16.
Proof.[Corollary 18] Let Π≁(n+1) denote those permutations π′ ∈ Πn+1 such
that if Ranking uses π′ when the input is MonotoneG (with U = V = n + 1),
n+1 (the last item in π′) is not matched. By definition of a(n, i) the
then π′
expression (n+1)!−a(n+1, n+1) can be interpreted as Π≁(n+1). We describe a
bijection B between (Πn, [n+1]) and Πn+1. The bijection will have the property
that given a pair (π ∈ Πn, i ∈ [n + 1]) the resulting permutation B(π, i) ∈ Πn+1
belongs to Π≁(n+1) if and only if ui is matched, thus proving the Corollary.
We now describe the bijection for a given π ∈ Πn and i ∈ [n + 1]:
• B(π, n + 1): place vn+1 at location n + 1. This gives one permutation that
we call π→(n+1).
• B(π, i) for 1 ≤ i ≤ n: place vn+1 at location i and place vi at location
n + 1. This gives n additional permutations, named π↔i (the notation ↔
indicates that vn+1 is swapped with vi).
There are three cases to consider:
• i = n + 1. In π→(n+1) the item vn+1 at location n + 1 is matched, because
it is a neighbor of all vertices in U .
22
• ui ∈ U is matched in π. Then all vertices up to ui are also matched in
π↔i, and to items at locations no later than n. This is because the only
differences between π and π↔i involve vertices vi and vn+1, and both of
them are neighbors of all arriving vertices up to and including ui. None
of the vertices ui+1, . . . , un+1 is a neighbor of vi, hence in π↔i the item
vi ∈ V at location n + 1 is not matched.
• ui ∈ U is not matched in π. In this case ui will not be matched to any
of the first n items of π↔i (again, because the only differences between
π and π↔i involve vertices vi and vn+1, and both of them are neighbors
of all arriving vertices up to and including ui). Consequently, ui will be
matched to vi that is at location n + 1 in π↔i.
(cid:4)
23
|
Subsets and Splits
No saved queries yet
Save your SQL queries to embed, download, and access them later. Queries will appear here once saved.